Pulv2002NeuroscLan

333

Transcript of Pulv2002NeuroscLan

Page 1: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 1/331

Page 2: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 2/331

The Neuroscience of Language

On Brain Circuits of Words and Serial Order

A realistic model of language should specify the mechanisms under-

lying language use and comprehension. A neurobiological approach

has been shown to be an effective means toward this end. The Neu-roscience of Language provides results of brain activation studies,

patients with brain lesions, and hints from computer simulations of 

neural networks to help answer the question: How is language orga-

nized in the human brain?

At the book’s core are neuronal mechanisms that is, the nerve cell

wiring of language in the brain. Neuronal models of word and serial-

order processing are presented in the form of a computational and

connectionist neural network. The linguistic emphasis is on wordsand elementary syntactic rules. The book introduces basic knowledge

from disciplines relevant in the cognitive neuroscience of language.

Introductory chapters focus on neuronal structure and function, cog-

nitive brain processes, the basics of classical aphasia research and

modern neuroimaging of language, neural network approaches to lan-

guage, and the basics of syntactic theories. The essence of the work

is contained in chapters on neural algorithms and networks, basic

syntax, serial-order mechanisms, and neuronal grammar. Through-

out, excursuses illustrate the functioning of brain models of language,

some of which are simulations accessible as animations on the book’s

accompanying web site.

This self-contained text and reference puts forth the first system-

atic model of language at a neuronal level that is attractive to lan-

guage theorists but that is also well grounded in empirical research.

The Neuroscience of Language bridges the gap between linguistics

and brain science, appealing to advanced students and researchers in

neuroscience, linguistics, and computational modeling.

Friedemann Pulvermuller is a senior scientist at the Medical Research

Council Cognition and Brain Sciences Unit in Cambridge.

Page 3: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 3/331

Page 4: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 4/331

The Neuroscience of Language

On Brain Circuits of Wordsand Serial Order

FRIEDEMANN PULVERM ¨ ULLERCognition and Brain Sciences UnitMedical Research Council

Page 5: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 5/331

Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São Paulo

Cambridge University PressThe Edinburgh Building, Cambridge , United Kingdom

First published in print format

isbn-13 978-0-521-79026-0 hardback 

isbn-13 978-0-521-79374-2 paperback 

isbn-13 978-0-511-06708-2 eBook (NetLibrary)

© Cambridge University Press 2002

2003

Information on this title: www.cambridge.org/9780521790260

This book is in copyright. Subject to statutory exception and to the provision of relevant collective licensing agreements, no reproduction of any part may take place

 without the written permission of Cambridge University Press.

isbn-10 0-511-06708-9 eBook (NetLibrary)

isbn-10 0-521-79026-3 hardback 

isbn-10 0-521-79374-2 paperback 

Cambridge University Press has no responsibility for the persistence or accuracy of s for external or third-party internet websites referred to in this book, and does notguarantee that any content on such websites is, or will remain, accurate or appropriate.

Published in the United States of America by Cambridge University Press, New York 

 www.cambridge.org

-

-

-

-

-

-

Page 6: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 6/331

Contents

Preface page xi

1 A Guide to the Book 1

1.1 Structure and Function of the Book 1

1.2 Paths Through the Book 2

1.3 Chapter Overview 41.3.1 Chapter 1: A Guide to the Book 41.3.2 Chapter 2: Neuronal Structure and Function 41.3.3 Chapter 3: From Aphasia Research to

Neuroimaging 41.3.4 Chapter 4: Words in the Brain 41.3.5 Excursus E1: Explaining Neuropsychological

Double Dissociations 51.3.6 Chapter 5: Regulation, Overlap, and Web Tails 5

1.3.7 Chapter 6: Neural Algorithms and NeuralNetworks 5

1.3.8 Chapter 7: Basic Syntax 51.3.9 Chapter 8: Synfire Chains as the Basis of Serial

Order in the Brain 61.3.10 Chapter 9: Sequence Detectors 61.3.11 Chapter 10: Neuronal Grammar 61.3.12 Chapter 11: Neuronal Grammar and

Algorithms 6

1.3.13 Excursus E2: Basic Bits of Neuronal Grammar 71.3.14 Excursus E3: A Web Response to a Sentence 71.3.15 Chapter 12: Refining Neuronal Grammar 71.3.16 Excursus E4: Multiple Reverberation for

Resolving Lexical Ambiguity 71.3.17 Excursus E5: Multiple Reverberations and

Multiple Center Embeddings 7

v

Page 7: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 7/331

vi Contents

1.3.18 Chapter 13: Neurophysiology of Syntax 81.3.19 Chapter 14: Linguistics and the Brain 8

2 Neuronal Structure and Function 9

2.1 Neuronal Structure 10

2.1.1 Anatomy of a Nerve Cell 102.1.2 Basics of the Cortex 132.1.3 Internal Wiring of the Cortex 16

2.2 Neuronal Function and Learning 18

2.3 Principles and Implications 20

2.4 Functional Webs in the Cortex 22

2.4.1 Why Numerous Neurons Should Cooperate 222.4.2 The Need for Connecting Neurons in Distant

Cortical Areas 232.5 Defining Functional Webs 24

2.6 Evidence for Functional Webs 26

2.7 A View of Cortical Function 28

2.8 Temporal Dynamics in Functional Webs: Ignition

and Reverberation 29

3 From Classic Aphasia Research to ModernNeuroimaging 33

3.1 Aphasiology 33

3.2 Laterality of Language 39

3.3 Neuroimaging of Language 44

3.4 Summary 48

4 Words in the Brain 50

4.1 Word-Form Webs 50

4.2 Category-Specific Word Webs 56

4.2.1 Visually Related and Action Words 564.2.2 Sub-types of Action Words 62

4.3 The Time Course of Lexical and Semantic Activation 62

4.4 Summary and Conclusions 64

Excursus E1: Explaining Neuropsychological DoubleDissociations 66

E1.1 Functional Changes in a Lesioned Network: The

Nonlinear Deterioration of Performance with

Growing Lesion Size 67

E1.2 Broca’s vs. Wernicke’s Aphasias 69

Page 8: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 8/331

Contents vii

5 Regulation, Overlap, and Web Tails 74

5.1 Regulation of Cortical Activity 75

5.1.1 The Overactivation Problem in AutoassociativeMemories 755.1.2 Coincidence vs. Correlation 765.1.3 Sparse Coding 765.1.4 A Cybernetic Model of Feedback Regulation

of Cortical Activity 785.1.5 Striatal Regulation of Cortical Activity 805.1.6 Summary 81

5.2 Overlapping Representations 82

5.2.1 Homophones and Form-Related Words 835.2.2 Synonyms 875.2.3 Prototypes and Family Resemblance 88

5.3 Web Tails 91

5.3.1 Affective and Emotional Meaning 915.3.2 Linking Phonological, Orthographical, and

Meaning-Related Information 915.4 Summary 95

6 Neural Algorithms and Neural Networks 966.1 McCulloch and Pitts’s Logical Calculus as a

Starting Point 97

6.2 Symbolic Connectionist Models of Language 107

6.3 Distributed Connectionist Models of Language 112

6.4 Hot Topics in Neural Network Research on Language 115

6.4.1 Word Category Deficits 1156.4.2 The Development of Rules in the Brain 119

7 Basic Syntax 124

7.1 Rewriting Rules 125

7.2 Center Embedding 128

7.3 Discontinuous Constituents and Distributed Words 134

7.4 Defining Word Categories in Terms of Complements:

Dependency Syntax 139

7.5 Syntactic Trees 141

7.6 Questions for a Neuronal Grammar 144

8 Synfire Chains as the Basis of Serial Order in the Brain 147

8.1 Neurophysiological Evidence and Neuronal Models 147

8.2 A Putative Basis of Phonological Processing 151

8.3 Can Synfire Chains Realize Grammar? 154

8.4 Functional Webs Composed of Synfire Chains 156

8.5 Summary 158

Page 9: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 9/331

viii Contents

9 Sequence Detectors 159

9.1 Movement Detection 159

9.2 Sequence Detectors for Word Strings 161

9.3 Sequence Detectors and Syntactic Structure 163

10 Neuronal Grammar 168

10.1 The Story So Far 169

10.2 Neuronal Sets 169

10.3 Threshold Control 175

10.4 Sequence Detection in Networks of Neuronal Sets 177

10.5 Activity Dynamics of Sequence Detection 181

10.6 Lexical Categories Represented in Neuronal Sets 18610.6.1 Why Lexical Categories? 18610.6.2 Lexical Ambiguity 18710.6.3 Lexical Categories as Sets of Sequence Sets 19010.6.4 Neuronal Requirements of a Grammar Machine 19210.6.5 Lexical Disambiguation by Sequence Sets 194

10.7 Summary: Principles of Neuronal Grammar 200

11 Neuronal Grammar and Algorithms 207

11.1 Regular Associations, Associative Rules 20711.2 A Formalism for Grammar Networks 209

11.3 Some Differences Between Abstract and

Neuronal Grammar 211

11.4 Summary 214

Excursus E2: Basic Bits of Neuronal Grammar 215

E2.1 Examples, Algorithms, and Networks 215

E2.2 Grammar Circuits at Work 217

E2.2.1 Simulation 1: Acceptance of a Congruent String 219E2.2.2 Simulation 2: Processing of an Incongruent String 220E2.2.3 Simulation 3: Processing of a Partly Congruent

String 222

Excursus E3: A Web Response to a Sentence 224

E3.1 The Grammar Algorithm and Network 225

E3.2 Sentence Processing in Syntactic Circuits 227

E3.2.1 Global Characteristics 228E3.2.2 Specific Features 231

E3.3 Discussion of the Implementation and Simulation 232

12 Refining Neuronal Grammar 235

12.1 Complements and Adjuncts 236

12.2 Multiple Activity States of a Neuronal Set 238

12.2.1 The Concept of Multiple Reverberation 239

Page 10: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 10/331

Contents ix 

12.2.2 Some Revised Principles of Neuronal Grammar 24112.3 Multiple Center Embedding 245

12.4 Summary, Open Questions, and Outlook 247

Excursus E4: Multiple Reverberation for ResolvingLexical Ambiguity 250

Excursus E5: Multiple Reverberations and MultipleCenter Embeddings 255

13 Neurophysiology of Syntax 265

13.1 Making Predictions 265

13.2 Neuronal Grammar and Syntax-RelatedBrain Potentials 266

14 Linguistics and the Brain 270

14.1 Why Are Linguistic Theories Abstract? 270

14.2 How May Linguistic Theory Profit from a

Brain Basis? 272

References 277

 Abbreviations 297

 Author Index 301

Subject Index 307

Page 11: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 11/331

Page 12: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 12/331

Preface

How is language organized in the human brain? This book provides results of 

brain activation studies, facts from patients with brain lesions, and hints from

computer simulations of neural networks that help answer this question.

Greateffortwasspenttospellouttheputativeneurobiologicalbasisofwordsand sentences in terms of nerve cells, or neurons. The neuronal mechanisms –

that is, the nerve cell wiring of language in the brain – are actually in thefocus. This means that facts about the activation of cortical areas, about

the linguistic deficits following brain disease, and the outcome of neural

network simulations will always be related to neuronal circuits that could

explain them, or, at least, could be their concrete organic counterpart in thebrain.

In cognitive neuroscience, the following questions are commonly asked

with regard to various higher brain functions, or cognitive processes, includ-

ing language processes:(1) Where? In which areas of the brain is a particular process located?

(2) When? Before and after which other processes does the particularprocess occur?

(3) How? By which neuron circuit or which neuron network type is the

particular process realized?

(4) Why? On the basis of which biological or other principles is the par-

ticular process realized by this particular network, at this particular

point in time, and at these particular brain loci?

The ultimate answer to the question of language and the brain implies ans-

wers to these questions, with respect to all aspects of language processing.

The aspects of language relevant here include the physical properties of 

speech sounds and the sound structure of individual languages as specified

by phonetics and phonology, the meaning and use of words and larger units

 xi

Page 13: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 13/331

 xii Preface

of language as specified by semantics and pragmatics, and the rules under-

lying the serial ordering of meaningful language units in sentences or larger

sequences as specified by syntax. All of these aspects are addressed, althoughan emphasis is put on words and elementary syntactic rules.

Lesion studies and brain imaging studies have revealed much importantinformation relevant for answering “Where” questions of type (1), about the

relevant brain loci. Fast neurophysiological imaging could also shed light

on the temporal structure of language processes in the millisecond range,

thereby answering “When” questions of type (2). Type (3) “How” questions

about the underlying mechanisms are sometimes addressed in neural net-

work studies, but many ideas – in particular, ideas about syntactic structures –are still formulated in terms so abstract that it is difficult to see possible con-nections to the neuronal substrate of the postulated processes. A common

excuse is that we do not know enough about the brain to specify the mecha-

nisms of grammar, and of language in general, in terms of neurons. I never

found this to be a very good excuse. Only if we theorize about concrete

language circuits can we ever understand them. Therefore, this book putsforward concrete, or neuronal , models of word and serial-order processing.

The introduced models are linked to general neuroscientific principles, andare thereby used to answer aspects of “Why” questions of type (4) as well.

This book came about because my abstract theorizing about neuron cir-

cuits that could realize aspects of language was initiated by various discus-

sions with Valentino Braitenberg and Almut Schuz, with whom I had thepleasure of collaborating at the Max Planck Institute of Biological Cyber-

netics in Tubingen, Germany. This was before and at the very start of the

so-called Decade of the Brain, with much less data in hand than is available

today. However, when collecting brain imaging data myself in subsequentyears, I found it necessary to have a brain-based model of language as a

guideline for designing experiments. I also found the concepts and mecha-nisms developed and envisaged earlier quite helpful for the interpretation

and explanation of new results. In the context of neuroimaging studies, I

should mention Werner Lutzenberger, my most important teacher in the

neurophysiological domain. I thank Almut, Valentino, and Werner for the

countless discussions about theory and data that laid the foundation for this

book.However, many more colleagues and friends contributed substantially.

Robert Miller was a critical reader of many manuscripts preceding this text

and provided me with indispensable knowledge, particularly neuroanatomi-

cal in nature. Stefano Crespi-Reghizzi, an expert in automata theory and

computational approaches to language, was kind enough to teach me thebasics from these disciplines, thereby preventing me from serious mistakes.

Page 14: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 14/331

Preface xiii

Helmut Schnelle, a theoretical linguist and one of the few who seriously

attempt at connecting syntax and the brain, provided me with a very de-

tailed and equally helpful critique of an earlier version of this book. Com-

parison with his approach to the representation of serial order in the brain

helped me greatly when formulating putative grammar mechanisms. Detlef 

Heck checked some of the neurophysiology and neuroanatomy sections and

helped me avoid errors there. William Marslen-Wilson read the entire book

with the eyes of a psycholinguist, and provided me with probably the most

detailed critique of an earlier version of this book, thereby initiating sev-

eral improvements. I am indebted to them all for their substantial help. I

also thank numerous other colleagues who commented on text passages orideas, including Joe Bogen, Michel Caillieux, Thomas Elbert, Gerd Fritz,

Joaquin Fuster, Sarah Hawkins, Risto Ilmoniemi, Risto Naatanen, Dennis

Norris, Lee Osterhout, Gunther Palm, Brigitte Rockstroh, Arnold Scheibel,

John Schumann, and, of course, my closest colleagues, Ramin Assadollahi,

Olaf Hauk, Bettina Neininger, Yury Shtyrov, and, most importantly, my wife

Bettina Mohr, not only because she performed the first experiment that made

me feel that the kind of model put forward here might be on the right track,

but also because she continuously gave me the radical and respectless cri-tique that is a necessary condition for scientific progress. Finally, I apologize

to my son Johannes David for the untold goodnight stories and unplayed

games that fell victim to writing the present pages.

Friedemann Pulvermuller

Cambridge, July 2001

Page 15: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 15/331

Page 16: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 16/331

The Neuroscience of Language

On Brain Circuits of Words and Serial Order

Page 17: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 17/331

Page 18: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 18/331

CHAPTER ONE

 A Guide to the Book

The neuroscience of language is a multidisciplinary field. The reader’s pri-

mary interest may therefore lie in various classical disciplines, including psy-

chology, neuroscience, neurology, linguistics, computational modeling, or

even philosophy. Because readers with different backgrounds may be inter-ested in different parts of this book, Chapter 1, Section 1.3 gives an overview

of the book contents and the gist of each chapter. In Section 1.1, the generalstructure of the book is explained; subsequently, paths through the book

are recommended for readers with different backgrounds and interests in

Section 1.2.

1.1 Structure and Function of the Book

The fourteen chapters of this book are mainly designed to convey one single

message: It is a good idea to think about language in terms of brain mecha-nisms – to spell out language in the language of neurons, so to speak. Makingthis point is not a new proposal. One can find similar statements in classi-

cal writings; for example, in Freud’s monograph on aphasia (Freud, 1891)

and other publications by neurologists in the late nineteenth century, and, of 

course, in modern brain-theoretical and linguistic publications (Braitenberg,

1980; Mesulam, 1990; Schnelle, 1996a). However, a systematic model of lan-

guage at the level of neurons as to date is not available, at least, not an

approach that would be both grounded in empirical research while at thesame time attacking a wide range of complex linguistic phenomena.

Apart from the main message, this book puts forward two principle

proposals: First, that words are represented and processed in the brain

by strongly connected distributed neuron populations exhibiting specific

topographies. These neuron ensembles are called word webs. Second, that

1

Page 19: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 19/331

2 A Guide to the Book

grammar mechanisms in the brain can be thought of in terms of neuronal

assemblies whose activity specifically relates to the serial activation of pairs

of other neuron ensembles. These assemblies are calledsequence sets

. Theproposal about word webs is presented in Chapter 4 and the one about se-

quence sets in Chapter 10. One may therefore consider Chapters 4 and 10the core chapters of this book.

As it happens, new proposals elicit discussion, which, in turn, makes re-

finement of the original proposals desirable. The word web proposal is being

refined in Chapters 5 and 8, and the proposal on grammar mechanisms is

further developed in Chapters 11 and 12. As stressed in the Preface, several

colleagues contributed to the refinements offered. The evolution of someof the ideas is documented in a recent discussion in the journal The Be-

havioral and Brain Sciences (Pulvermuller, 1999b). Summaries of ideas put

forward here can be found in related review papers (Pulvermuller, 2001,

2002).

Apart from presenting the two main proposals, the book is designed to

give the reader an introduction to basic knowledge from disciplines relevantin the cognitive neuroscience of language. Chapter 2 offers an introduction

to neuroscience and cognitive brain processes. Chapter 3 introduces basicsabout classical aphasia research and modern neuroimaging of language. Two

more introductory chapters follow approximately in the middle of the book.

Chapter 6 features neural network approaches to language, and Chapter 7

introduces basics of syntactic theories. These introductory chapters werewritten to make the book “self-contained,” so that ideally speaking no prior

special knowledge would be required to understand it.

Interspersed between the chapters are five excursuses, labeled E1 through

E5, which illustrate the functioning of brain models of language. In eachexcursus, one or more simple simulations are summarized that address an

issue raised in the preceding chapter. Computer simulations of the main syn-dromes of aphasia (Excursus E1) are included along with simulations of the

processing of simple (Excursus E2) and gradually more complex (Excursuses

E3–E5) sentences in brain models of grammar. Some of the simulations are

available as animations accessable through the Internet.

1.2 Paths Through the Book

Clearly, the reader can choose to read through the book from beginningto end. However, because not all issues covered by the book may be in

the inner circle of one’s personal “hot topics,” it may be advantageous to

have available alternatives to this global strategy. One alternative would

be to take a glance at the main chapters (4 and 10) or at the introductory

chapter concerning the topic one is particularly keen on. However, one may

Page 20: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 20/331

1.2 Paths Through the Book 3

Table 1.1. Routes through the book recommended to readers primarilyinterested in neuroscience, linguistics, or neuronal modeling, respectively.

Chapter numbers and headings are indicated. Headings are sometimesabbreviated. Excursuses are referred to by the letter E plus a numberand by abbreviated headings. For further explanation, see text.

Neuroscience Route Linguistics Route Modeling Route

2 Neuronal structure 4 Words in the brain 4 Words in the brain

and function

3 Aphasia and 7 Basic syntax 5 Regulation, overlap,

neuroimaging web tails

4 Words in the brain 8 Synfire chains 6 Neural networks

E1 Double dissociations 9 Sequence detectors E1 Double dissociations

5 Regulation, overlap, 10 Neuronal grammar E2 Basic bits of neuronal

web tails grammar

8 Synfire chains 11 Neuronal grammar E3 Web response to a

and algorithms sentence

9 Sequence detectors 12 Refining neuronal E4 Lexical ambiguity

grammar

13 Neurophysiology of 14 Linguistics and the E5 Center embedding

syntax brain

wish to dive deeper into the matter while still primarily following one’s

interests.

For this latter purpose, three paths through the book are offered for a

reader primarily interested in neuroscience, linguistics, and neurocomputa-

tional modeling. If one chooses one of these options, one should be awarethat the routes are not self-contained and consultation of other chapters

may be relevant occasionally. To facilitate detection of relevant information

in other chapters of the book, multiple cross-references have been added

throughout.

The three paths through the book are presented in Table 1.1. Please con-

sult the overview, Section 1.3, for details about chapter contents.

It is difficult to decide what to recommend to a reader primarily interested

in psychology. Because psychology is a rather wide field, the best recom-mendation may depend primarily on the subdiscipline of interest. Readers

interested in neuropsychology and psychophysiology can be recommended

to follow the neuroscience route, whereas those interested in cognitive psy-

chology may tend more toward modeling aspects. The neuroscience route

would also be recommended to the reader focusing on neuroimaging or

neurology. A philosopher may be most interested in the open questions that

accumulate in Chapters 5, 12, and 14.

Page 21: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 21/331

4 A Guide to the Book

1.3 Chapter Overview 

1.3.1 Chapter 1: A Guide to the BookThe main purpose of the book and its structure are explained briefly. Recom-mendations are given concerning how to use the book if one is interested

primarily in its neuroscience, linguistics, or modeling aspects. The gist of each

book chapter is summarized briefly.

1.3.2 Chapter 2: Neuronal Structure and Function

Chapter 2 introduces basics about the anatomy and function of the neuronand the cortex. Principles of cortical structure and function are proposed

that may be used as a guideline in cognitive brain research. The concept

of a distributed functional system of nerve cells, called functional web, isintroduced and discussed in the light of neurophysiological evidence.

1.3.3 Chapter 3: From Aphasia Research to Neuroimaging

Basics about aphasias, language disorders caused by disease of the adult

brain, are summarized. Aphasia types and possibilities on explaining some

of their aspects are being discussed. The issue of the laterality of language to

the dominant hemisphere – usually the left hemisphere – is mentioned, and

theories of laterality and interhemispheric interaction are covered. Basic

insights in the functional architecture of the cortex as revealed by modernneuroimaging techniques are also in the focus. The conclusion is that some,

but not all, insights from classical aphasia research about the localizationof cortical language functions can be confirmed by neuroimaging research.

However, language processes seem to be much more widely distributed than

previously assumed. The question about the cortical locus of word semantics,

as such, has found contradicting answers in recent imaging research.

1.3.4 Chapter 4: Words in the Brain

The proposal that words are cortically represented and processed by dis-tributed functional webs of neurons is elaborated and discussed on the basis

of recent neuroimaging studies. The data support the postulate that words

and concepts are laid down cortically as distributed neuron webs with dif-

ferent topographies. The strongly connected distributed neuron ensembles

representing words are labeled word webs. Word webs may consist of a

phonological part (mainly housed in the language areas) and a semantic part

Page 22: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 22/331

1.3 Chapter Overview 5

(involving other areas as well). For example, processing of words with strong

associations to actions and that of words with strong visual associations ap-

pears to activate distinct sets of brain areas. Also, different subcategoriesof action words have been found to elicit differential brain responses. Thissupports the proposed model.

1.3.5 Excursus E1: Explaining NeuropsychologicalDouble Dissociations

A simulation is presented that allows for the explanation of neuropsycho-

logical double dissociations on the basis of distributed functional webs of neurons. The nonlinear decline of performance of the models with lesion

size and its putative neurological relevance are also mentioned.

1.3.6 Chapter 5: Regulation, Overlap, and Web Tails

Chapter 5 deals with open issues remaining from earlier chapters. How could

a regulation device controlling activity in the cortex be organized? How

would words with similar meaning but different form, or words with simi-lar form but different meaning, be realized in the brain? Would the brain’s

word processor be restricted to the cortex, or can word webs have subcortical

“tails”? One postulate is that multiple overlap between cortical representa-tions exists between word representations.

1.3.7 Chapter 6: Neural Algorithms and Neural Networks

An introduction into neural network models is given. McCulloch and Pitt’s

theory is sketched and perceptron-based simulations are featured. Symbolic

connectionist approaches are also discussed briefly. Among the hot topics

featured are the explanation of word category deficits as seen in neurological

patients and the development of rules in infants’ brains.

1.3.8 Chapter 7: Basic Syntax

A few terms and theoretical approaches to syntax are introduced. Phrase

structure grammars, dependency grammars, and more modern proposals

rooted in these classic approaches are discussed. Syntactic problems such

as those associated with long-distance dependencies and center embeddings

are mentioned. Chapter 7 ends with a list of issues with which grammar

circuits should cope.

Page 23: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 23/331

6 A Guide to the Book

1.3.9 Chapter 8: Synfire Chains as the Basis of Serial Orderin the Brain

One type of serial-order mechanism in the brain for which there is evidencefrom neurophysiological research is featured. Called a synfire chain, it con-

sists of local groups of cortical neurons connected in sequence, with loops

also allowed for (reverberatory synfire chain). The synfire model of serial or-der is found to be useful in modeling phonological–phonetic processes. It is

argued, however, that a synfire model of syntax does not appear to be fruitful.

1.3.10 Chapter 9: Sequence DetectorsA second type of serial-order mechanism exists for which there is evidence

from brain research. It is the detection of a sequence of neuron activationsby a third neuronal element called the sequence detector . The evidence for

sequence detectors comes from various brain structures in various creatures.

It is argued that sequence detectors may operate on sequences of activations

of word webs, and that these may be part of the grammar machinery in the

brain.

1.3.11 Chapter 10: Neuronal Grammar

Neuronal sets are defined as functional webs with four possible activity states:

inactivity (O), full activation or ignition (I), sustained activity or reverber-

ation (R) and neighbor-induced preactivity or priming (P). Reverberation

and priming levels can vary. Grammar networks are proposed to be made

up of two types of neuronal sets: word webs and sequence sets. Sequencesets respond specifically to word sequences. The lexical category of wordsand morphemes is represented by a set of sequence sets connected directly

to word webs. Words that can be classified as members of different lexical

categories have several mutually exclusive sets of sequence sets. Activity dy-

namics in the network are defined by a set of principles. A grammar network,

also called neuronal grammar , can accept strings of words or morphemes oc-

curring in the input, including sentences with long-distance dependencies.

The hierarchical relationship between sentence parts becomes visible in theactivation and deactivation sequence caused by an input string.

1.3.12 Chapter 11: Neuronal Grammar and Algorithms

Three types of formulas are introduced that describe a neuronal grammar

network:

Page 24: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 24/331

1.3 Chapter Overview 7

1. Assignment formulas are definitions of connections between input

units and lexical category representations and are analogous to lexicon

or assignment rules of traditional grammars.

2. Valence formulas are definitions of lexical categories in terms of se-

quencing units and have some similarity to dependency rules included

in dependency grammars.

3. Sequence formulas are definitions of connections between sequencing

units and have no obvious counterpart in traditional grammars.

1.3.13 Excursus E2: Basic Bits of Neuronal Grammar

Simple word strings are discussed on the basis of grammar networks com-

posed of sequence sets and word webs. How the network accepts a string

and how the network behaves if it fails to do so is discussed.

1.3.14 Excursus E3: A Web Response to a Sentence

Processing of an ordinary sentence is simulated in a neuronal grammar

architecture. The sentence exhibits six morphemes, subject–verb agreement,a distributed word, and other interesting properties.

1.3.15 Chapter 12: Refining Neuronal Grammar

A revision of the grammar model is proposed that requires stronger as-

sumptions. The core assumption is that neuronal sets exhibit multiple states

of reverberation and priming. In the new architecture, the relationship be-

tween words and lexical categories is now dynamic.

1.3.16 Excursus E4: Multiple Reverberation for ResolvingLexical Ambiguity

Implementation of multiple lexical category representations of words using

mutually exclusive sets of sequence sets allows for modeling sentences in

which the same word form is being used twice, as a member of different

lexical categories.

1.3.17 Excursus E5: Multiple Reverberations and MultipleCenter Embeddings

A network with dynamic binding between word and lexical category repre-

sentations and the option to activate each neuronal set is introduced on the

Page 25: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 25/331

8 A Guide to the Book

background of the machinery discussed in Chapters 10 and 11. This more

advanced architecture now models the processing of grammatically complex

sentences that include center embeddings.

1.3.18 Chapter 13: Neurophysiology of Syntax

Grammatically incorrect “sentences” elicit specific physiological brain re-

sponses. Two such physiological indicators of grammatical deviance are dis-

cussed. The neuronal grammar proposal is related to these data, and a pu-

tative neurobiological explanation for them is offered.

1.3.19 Chapter 14: Linguistics and the Brain

Linguistics and brain science must merge. This is reemphasized in Chapter 14

where putative advantages of a brain-based language theory are highlighted.

Page 26: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 26/331

CHAPTER TWO

Neuronal Structure and Function

A realistic model of language must specify the mechanisms underlying lan-

guage use and comprehension. What are the relevant mechanisms? It is

certain that it is the human brain that provides the mechanisms realizing

language, and it is almost equally certain that language mechanisms are or-ganized as nerve cells and their mutual connections. A realistic model of 

language, therefore, must specify the putative organic basis of language useand language comprehension in terms of neurons, neuronal connections,

and neuron circuits. This does not necessarily mean that the model must

specify each and every single neuron that participates, but it does mean that

the circuits believed to underlie language function should be specified as faras possible and relevant. Rather than saying that a language sound, word,

or syntactic rule is represented in the brain, period, one may wish to learn

in which way such sounds, words, or rules are laid down. Therefore, it is

necessary to introduce known neuronal mechanisms and others that can beinferred from more recent data.Chapter 2 gives a brief introduction to neuronal mechanisms, with special

emphasis on those mechanisms that may be relevant for organizing lan-

guage. The chapter first addresses questions about neuronal architecture or

structure (Section 2.1). What is a nerve cell or neuron? What is the global

structure of the cortex, the portion of the brain most important for language?

The idea here is that these anatomic structures are related to the computa-

tions in which they are involved. The brain machinery is not just one arbitraryway of implementing the processes it realizes, as, for example, any hardware

computer configuration can realize almost any computer program or piece

of software. The claim is that, instead, the hardware reveals aspects of the

program. Neuronal structure is information (Braitenberg, 1971). In other

words, it may be that the neuronal structures themselves teach us about as-pects of the computational processes that are laid down in these structures.

9

Page 27: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 27/331

10 Neuronal Structure and Function

The paragraphs on cortical structure detail a few structural features and

elaborate on their functions as well.

As a next step, additional functional properties of the neuron are high-lighted – in particular, the question of how information may be stored in

nerve cells and their connections is addressed (Section 2.2). The structuraland functional properties of the neuron and the cortex are summarized by

three conclusions that are used later in the book as axioms or principles for

theorizing about language mechanisms (Section 2.3). A few thoughts follow

about how neurons in the cortex may interact to yield what is sometimes

called the cognitive or higher brain functions (Section 2.4). These terms can

refer to language-related processes, of course, but to other complex percep-tual and action-related processes as well. Finally, the concept of a functional 

web is introduced and grounded in empirical evidence.

2.1 Neuronal Structure

2.1.1 Anatomy of a Nerve Cell

The functional element of the brain and the nervous system is the nervecell, orneuron. Figure 2.1 shows an example of one type of neuron, so-called pyramidal cells. This is the most common neuron type in the largest structure

of the human brain, the cortex. Like most other neuron types, a pyramidal

cell consists of dendrites, a cell body,andan axon. In Figure 2.1, the cell body

is the thick speck in the middle. The thicker lines departing from the cellbody and running sideways and upward are the dendrites, and the thin line

running downward is the axon. The dendrites and the axon branch multiply.

The neuron receives signals from other neurons through its dendrites, andit transmits its own signals to other neurons through its axon and its many

branches. Whereas the dendrites are short and hardly reach loci 1 mm away

from the cell body, many pyramidal neurons have very long axon branches,in addition to the short axon branches shown in the figure. The long axon

branch of a neuron can be 10 cm or more long, and its synapses can contact

other neurons in distant cortical areas or subcortical nuclei.

Signals are passed between neurons by contact buttons called synapses.

The synapses delivering incoming – or afferent – signals to the neuron arelocated mainly on the dendrites and a few sit on the cell body. The dendritesof pyramidal cells appear to be relatively thick in Figure 2.1 because they

carry many hillocks, called spines, on which most of the afferent synapses are

located (Fig. 2.2). The spine’s size and shape may be related to the functional

connection between two pyramidal neurons. Outgoing – or efferent – signals

are transmitted through synapses at the very ends of the axon’s branches.

Page 28: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 28/331

    a    r    e    a

    a    r    e    a

      A   -    s    y    s     t    e    m

    a    p

      i    c    a      l d

    e    n      d    r      i     t    e    s

      b    a

    s    a      l

    c    o      l      l    a     t    e    r    a      l

    a    x    o    n

      B

   -    s    y    s     t    e    m

      A   -    s    y    s     t    e    m

       F       i     g     u     r     e

       2  .

       1  .     (     L    e     f    t     )     G    o     l    g     i  -    s    t    a     i    n    e     d    p   y    r    a    m     i     d    a     l    c    e     l     l .     T

     h    e    c    e     l     l     b    o     d   y     i    s     i    n    t     h    e    m     i     d     d     l    e

    a    n     d    t     h    e    a    p     i    c    a     l    a    n     d     b    a    s    a     l     d

    e    n     d    r     i    t    e    s    c    a    n     b    e     d     i    s    t     i    n    g   u     i    s     h    e     d    a     l    o    n    g

   w     i    t     h    t     h    e    a   x    o    n    a    n     d     i    t    s     l    o    c    a     l     b    r    a    n    c     h    e    s .

     T     h    e     l    o    n    g    a   x    o    n     b    r    a    n    c     h    r   u    n    n     i    n    g     i    n    t    o    t     h    e   w     h     i    t    e    m    a    t    t    e    r    a    n     d    c    o    n    n    e    c    t     i    n    g

   w     i    t     h    n    e   u    r    o    n    s     i    n    o    t     h    e    r     b    r    a     i    n    a    r    e    a    s     i    s

    n    o    t    s     h    o   w    n .

     R    e    p    r     i    n    t    e     d

   w     i    t     h    p    e    r    m     i    s    s     i    o    n     f    r    o    m

     B    r    a     i    t    e    n     b    e    r    g ,

     V . ,

     &     S    c     h     ¨   u   z ,

     A .

     (     1     9     9     8     ) .     C   o   r    t   e   x   :   s    t   a    t     i   s    t     i   c   s   a   n     d   g   e   o   m

   e    t   r   y   o     f   n   e   u   r   o   n   a     l   c   o   n   n   e   c    t     i   v     i    t   y     (     2    e     d .     ) .

     B    e    r     l     i    n    :     S    p    r     i    n    g    e    r     (     R     i    g     h    t

     )     T     h    e    c    o    r    t     i    c    a     l   w     i    r     i    n    g    t     h    r    o   u    g     h

    t     h    e     l    o    n    g  -     d     i    s    t    a    n    c    e     A  -    s   y    s    t    e    m

    a    n     d     l    o    c    a     l     B  -    s   y    s    t    e    m

    o     f    c    o    n    n    e

    c    t     i    o    n    s     i    s     i     l     l   u    s    t    r    a    t    e     d .

     (     F    o    r    e   x    p

     l    a    n    a    t     i    o    n ,

    s    e    e    t    e   x    t .     )     F    r    o    m

     B    r    a     i    t    e    n     b    e    r    g ,

     V .

     (     1     9     9     1     ) .     L    o    g     i    c    s    a    t     d     i     f     f    e    r    e    n    t     l    e   v    e     l    s     i    n    t     h    e     b    r    a     i    n .     I    n

     I .     C     l   u    e    r     (     E     d .     ) ,

     A    t    t     i     d   e     l     C   o   n   g

   r   e   s   s   o     “     N   u   o   v     i   p   r   o     b     l   e   m     i     d   e     l     l   a

     l   o   g     i   c   a   e

     d   e     l     l   a     fi     l   o   s   o     fi   a     d   e     l     l   a   s     i   e

   n   z   a ,     ”

     V   o     l .     1     (    p    p .

     5     3  –     9     ) .     B    o     l    o

    g    n    a    :     F    r    a    n    c    o     A    n    g    e     l     i     E     d     i    t    o    r    e .

11

Page 29: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 29/331

12 Neuronal Structure and Function

Figure 2.2. (Left) A dendrite with spines covering its surface. (Right) Microphotographsof spines with different calibers. Most synapses between pyramidal cells are located onspines. Their shape influences the efficacy of signal transmission between the cells. FromBraitenberg, V., & Sch ¨ uz, A. (1998). Cortex: statistics and geometry of neuronal connectivity (2 ed.). Berlin: Springer.

A pyramidal neuron in the human cortex has about 5 × 10,000 (104) incom-ing synapses and about the same number of efferent synapses (DeFelipe &Farinas, 1992).

As mentioned, the dendrites of pyramidal cells branch multiply, resultingin a so-called dendritic tree. This treelike structure differs from the shape

of a typical conifer. It has sideways branches, called basal dendrites, and

in many cases one long upward branch, called the apical dendrite. Thisanatomical subdivision in apical and basal dendrites has functional mean-

ing (Braitenberg, 1978b; Braitenberg & Schuz, 1998). Neurons located in

close vicinity appear to contact each other primarily through synapses on

their basal dendrites. Apical dendrites in upper cortical layers carry synapses

contacting distant neurons located in subcortical structures or in other ar-eas of the cortex. A similar subdivision has been proposed for the axon’sbranches. Some branches are short and contact neighboring neurons. How-

ever, the long axon reaching distant cortical areas has many additional

branches at its end and contacts other neurons there, usually on their apical

dendrites.

Thus, already the microanatomy of the pyramidal neuron suggests a

subdivision of cortical connections into a local and a long-distance system.

The local system is between basal dendrites and local axon collaterals. Be-cause it is wired through basal dendrites, it has been dubbed the B-system.

The long-distance system is between long axons and the apical dendrites

they contact and is therefore called the A-system. About half of a pyramidal

neuron’s synapses are in the short-distance B-system, and the other half in

the long-range A-system. It has been speculated that the local B-system and

Page 30: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 30/331

2.1 Neuronal Structure 13

the wide-range A-system have different functions and play different roles in

cognitive processing (Braitenberg, 1978b).

2.1.2 Basics of the Cortex

The part of the brain most relevant for language is the cerebral cortex.

This fact has been proved by neurological observations, in particular, the

fact that lesions in certain areas of the cortex lead to neurological language

impairment, aphasia, and by more recent imaging studies. The cortex consists

of two halves called the cortical hemispheres, one of which is usually more

important for language than the other (Broca, 1861).The cortex is an amply folded layer and, therefore, much of its surface is

not visible when looking at it from outside, but rather buried in the many

valleys or sulci. The cortical parts visible from the outside are the gyri. In

humans, the cortex is only 2- to 5-mm thick. This thin layer of densely packed

neurons rests on a fundament of white matter. The white matter has its color

from the isolating sheaths covering most of the A-system axons of corticalcells. Compared to the shining white of the white matter, the cortex itself 

looks gray because of the many gray cell bodies it contains. The white matterfundament of the cortex is much more voluminous than the gray matter. This

mere anatomical fact – the thin gray matter layer of cell bodies sitting on a

massive fundament of cables connecting the cell bodies – suggests that the

information exchange between cortical neurons that lie far apart may be animportant function of this structure.

The human cortex includes 1010 to 1011 neurons, about 85 percent of 

which are pyramidal cells (Braitenberg & Schuz, 1998). Because each of 

these neurons has about 104

synapses, the total number of cortical synapsesis in the order of 1014 or even higher. Apart from pyramidal cells, thereare other cell types. Some of these have inhibitory function. These neurons

are much smaller than pyramidal cells, and make only local connections

and receive input from many adjacent pyramidal cells. The small inhibitory

cells can dampen down cortical activity in case too many of their adjacent

excitatory pyramidal cells become active at one time.

Each hemisphere can be subdivided into some 50 to 100 areas. Figure 2.3

shows a lateral view of the left hemisphere with its most common area sub-division. These areas have been proposed by Brodmann (1909) and reflect

neuroanatomical properties of the cortical gray matter.

A cruder subdivision of the cortex is in terms of lobes. Figure 2.3 shows the

frontal lobe on its upper left, the parietal lobe at the top, the occipital lobe on

the right, and the temporal lobe at the bottom. Two major landmarks indicate

Page 31: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 31/331

14 Neuronal Structure and Function

Figure 2.3. Brodmann’s map of the human cortex. A lateral view on the left hemisphereis shown and the area numbers are indicated. Solid lines indicate area numbers and bro-ken lines indicate boundaries between areas. The shaded fields are primary areas. FromBrodmann, K. (1909). Vergleichende Lokalisationslehre der Groβhirnrinde. Leipzig: Barth.

boundaries between lobes. The central sulcus,orfissure,indicatedbythebold

line between Brodmann areas 3 and 4, is the boundary between the frontal

and the parietal lobes, and the Sylvian fissure, running horizontally belowareas 44 and 45 and above area 22, separates the temporal lobe from both

frontal and parietal lobes. The Sylvian fissure is important as a landmark,

because all of the areas most relevant for language are located in its close

vicinity. The areas next to the Sylvian fissure are called the perisylvian areas.

The occipital lobe at the back of the cortex is less easy to define based ongross anatomical landmarks. Gyri and sulci are sometimes referred to byspecifying their locus within a lobe. In the temporal lobe, for example, an

upper (superior), a middle, and a lower (inferior) gyrus can be distinguished,

and the same is possible in the frontal lobe, where, for example, the inferior

gyrus includes Brodmann areas 44 and 45.

When voluntary movements or actions are being performed, the neces-

sary muscle contractions are caused by activity in the cortex. When stim-

uli cause perceptual experiences, there is also activity in the cortex. Theefferent fibers through which the cortex controls muscle activity and the

afferent fibers transmitting information from the sensory organs to the cor-

tex originate from defined areas. These are the primary and secondary

areas. The primary areas most relevant for language processing are as

follows:

Page 32: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 32/331

2.1 Neuronal Structure 15

   r Brodmann area 17 in the posterior occipital lobe, where the fibers of the

visual pathway reach the cortex (primary visual cortex)   r

Brodmann area 41 in the superior temporal lobe, where the fibers of theauditory pathway reach the cortex (primary auditory cortex)   r Brodmann areas 1–3 in the postcentral gyrus of the parietal lobe, where

somatosensory input reaches the cortex (primary somatosensory cortex)   r Brodmann area 4 in the precentral gyrus of the frontal lobes, whose ef-

ferent neurons contact motor neurons and control muscle contractions

(primary motor cortex).

These sensory and motor fields are the shaded areas in Figure 2.3. Also

areas adjacent to the primary areas include efferent or afferent cortical con-

nections (e.g., He, Dum, & Strick, 1993). There are additional sensory path-ways for olfactory and gustatory input that project to brain areas not shown

in the diagram.

Each of the motor and sensory cortical systems is characterized bytopographical order of projections, meaning that adjacent cells in the sen-

sory organs project to adjacent cortical neurons, and adjacent body muscles

are controlled by adjacent neurons in the motor cortex.The somatotopy of the primary motor cortex is illustrated in Figure 2.4

(Penfield & Roberts, 1959). In this diagram, a frontal section of the cor-

tex along the precentral gyrus is schematized, and the body parts whose

movement can be caused by electrical stimulation of the respective part of 

the cortex is illustrated by the homunculus picture. The representations of 

the articulators – for example, tongue, mouth, and lips – are adjacent to eachother and lie at the very bottom of the illustration. Superior to them, the hand

representation is indicated, and the representation of the lower half of the

body, including the feet and legs, is further up and in the medial part of the

Figure 2.4. Penfield and Rassmussen (1950) in-vestigated the organization of the human mo-tor cortex by stimulating the cortical surface withweak currents and by recording which body mus-cles became active as a consequence of stimu-lation. The body muscles whose contraction wascaused by stimulation of these areas are indicatedby the homunculus picture. The map is consis-tent with the neuroanatomical projections of themotor cortex. It illustrates the somatotopic orga-nization of the primary motor cortex. FromPenfield, W., & Rassmussen, T. (1950). The cere-

bral cortex of man. New York: Macmillan.

Page 33: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 33/331

16 Neuronal Structure and Function

cortex. There are topographically ordered projections between body mus-

cles and loci of the primary motor cortex, and topographically ordered pro-

 jections exist in the visual, auditory, and somatosensory cortical systems aswell.

These projections are established early in life. However, they can be al-tered, for example, as a consequence of brain injuries. Research on cortical

reorganization has shown that the cortical representations can change dra-

matically, for example, as a consequence of deprivation (Buonomano &

Merzenich, 1998; Kujala, Alho, & Naatanen, 2000; Merzenich et al., 1983).

A change of the cortical areas involved in processing a particular input can

even be a consequence of learning. For example, string players and braillereaders show an altered function of somatosensory areas with enlarged cor-

tical representations of the extremities involved in complex sensorimotor

skills (Elbert et al., 1995; Sterr et al., 1998). This shows that the topographic

projections are not fixed genetically, but may vary within certain boundaries.

Because sensory projections to the cortex are being altered following depri-

vation and the acquisition of special skills, the projection map in Figure 2.4 isprobably appropriate only for nondeprived individuals without special skills.

Nevertheless, even after substantial cortical reorganization, the principle of topographical projections still appears to hold true (Merzenich et al., 1983).

2.1.3 Internal Wiring of the Cortex

Apart from the connections between the cortex and the outside world, itis relevant to say a few words about the internal wiring of the cortex. The

anatomy of the pyramidal cell makes it clear that there are different types of 

connections: short-range and long-range. As mentioned, the local links arethrough the B-system, that is, through connections between basal dendrites

and axon collaterals. The short-range connections are between neighbor-

ing neurons whose trees of dendrites and local axon collaterals overlap. Thelikelihood of such a contact between two neurons occupying almost the same

cortical space appears to be high. Estimates of the connection probabilities  p

of nearby neurons range between 10 percent and 80 percent, depending on

the method used (Braitenberg & Schuz, 1998; Hellwig, 2000; Mason, Nicoll,

& Stratford, 1991). Assuming a maximum of two connections, one in each di-rection, between any two neurons a group of n neurons can have a maximumof n(n− 1) direct connections between its members. The estimated num-

ber of connections between n nearby neurons, located, say, under the same

0.5 mm2 of cortical surface, would therefore be p× n(n− 1). This means

that 1,000 neurons selected by chance from a small piece of cortex would

have hundreds of thousands of mutual synaptic links – the exact number

Page 34: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 34/331

2.1 Neuronal Structure 17

being in the order of 100,000 to 800,000, depending on the actual parame-

ters chosen. This implies that these neurons can exert a strong influence on

each other, given they activate each other near synchronously. In essence,the high probability of local B-system links suggests local clusters of neurons

that exchange information intensely and therefore show similar functionalcharacteristics. Such local clusters, called columns, have been identified, for

example, in the visual system. A column includes neurons below 0.1–0.5 mm2

of cortical surface. The neurons in a column respond to similar visual stim-

ulus features and can thus be considered to be functionally related (Hubel,

1995). Although this functional relatedness is, in part, due to their common

sensory input, within-column connections are also likely to contribute.The long-distance A-system connections between pyramidal cells link

neurons in different cortical areas through their long axons and apical den-

drites. However, not every area is connected to every other area. The connec-

tion structure of the human cortex may be inferred from the results of neu-

roanatomical studies in animals, usually cats and macaque monkeys. These

animal studies provide evidence that most primary cortical areas do not con-tact each other through direct connections (Pandya & Yeterian, 1985). The

primary motor and somatosensory cortices, which lie adjacent to each other,represent the only exception. Adjacent areas, as a rule, are connected with

very high probability (> 70 percent; Young, Scannell, & Burns, 1995). For

pairs of distant areas, that is, areas with more than one other area between

them, this probability is lower (15–30 percent). Still, however, it is remark-able that, for example, in the macaque monkey, where nearly seventy areas

were distinguished, most areas would have links to ten or more distant areas

within the same cortical hemisphere. Even in the brain of the mouse, in which

only twelve local compartments were distinguished, each compartment wasfound to send out to and receive projections from five other compartments

on average (Braitenberg & Schuz, 1998). In addition, there are connectionsbetween most homotopic areas of the two hemispheres. Thus, long-distance

links directly connect many, although not all, cortical areas. Another im-

portant feature of corticocortical connectivity is that most between-area

connections are reciprocal; that is, they include neurons in both areas pro-

 jecting to the respective other area (Pandya & Yeterian, 1985; Young et al.,

1995).It is not certain that all of these properties of cortical connectivity inves-

tigated in monkeys and other higher mammals generalize to humans. A de-

tailed picture of cortical connectivity can be obtained only by using invasive

techniques that cannot be applied in humans for ethical reasons, although

important insights come from postmortem neuroanatomical studies inhumans (Galuske et al., 2000; Jacobs et al., 1993; Scheibel et al., 1985). In

Page 35: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 35/331

18 Neuronal Structure and Function

particular, conclusions on the pattern of long-distance connections of the

areas most important for language must be handled with care because these

areas do not have unique homologs in the monkey brain. However, a ten-tative generalization can be proposed in terms of the position of the areas

relative to primary areas.One such generalization is the following: The primary auditory cortex and

the primary motor cortex controlling the articulators are not linked directly.

Their connections are indirect, through inferior frontal areas anterior to the

mouth–motor cortex and superior temporal areas anterior, posterior, and

inferior to the primary auditory cortex (Deacon, 1992; Pandya & Yeterian,

1985). Therefore, if two neurons, one situated in the primary cortex control-ling mouth movements and the other in the piece of cortex receiving direct

input from fibers of the auditory pathway, want to communicate with each

other, they cannot do so directly. They must send their information to relay

neurons that link the two areas. Many such relay neurons are likely present

in the inferior frontal lobe (Brodmann’s areas 44 and 45) and in the superior

temporal lobe (Brodmann’s areas 22 and 42). A similar point made here forassociations between actions and auditory input can also be made for other

between-modality associations. Auditory–visual, somatosensory–visual, andvisual–action associations all require indirect links through relay neurons in

one or more nonprimary areas.

2.2 Neuronal Function and Learning

A neuron receives information, performs calculations on it, and sends out

information to other neurons. An afferent synapse on one of its dendrites can

produce an activating or inhibiting effect. If a signal is transmitted throughthe synapse, a so-called postsynaptic potential is elicited. There are excitatory

and inhibitory postsynaptic potentials (EPSPs and IPSPs). The postsynaptic

potential is propagated along the dendritic tree toward the cell body. Thesignal reaching the cell body is crucial for the neuron’s output. At the so-

called axon hillock, where the axon inserts from the cell body, computations

are being performed to the effect that if this signal is excitatory enough,

an output is created. If the potential increases rapidly or if it exceeds a

threshold value, a condition usually met when several incoming signals arrivesimultaneously, a new signal, called the action potential , is likely to be elicitedandthenpropagatedalongtheaxon.Theneuronissaidto fire when an action

potential is being created. The action potential is an all-or-none response –

either it is elicited or not, and does not vary in intensity. It propagates without

being attenuated from the cell body to the various efferent synapses at the

ends of the many branches of the axon. In contrast to the action potential

Page 36: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 36/331

2.2 Neuronal Function and Learning 19

with its all-or-none characteristics, the postsynaptic potential is an analog

signal that falls off gradually with time and distance along the dendrite.

Synaptic connections between cortical neurons are usually weak (Abeles,1991). Several simultaneous or near-simultaneous inputs causing strongly

overlapping postsynaptic potentials are therefore necessary to elicit an ac-tion potential in a neuron. However, the strength of a synaptic link can vary.

Synaptic links become stronger if the synapse is being used frequently. This is

the essence of a postulate put forward, among others, by Donald Hebb, who

said “that any two cells or systems of cells that are repeatedly active at the

same time will tend to become ‘associated,’ so that activity in one facilitates

activity in the other” (1949, p. 70). There is now strong evidence from single-and multiple-unit recordings proving that this postulate is correct (Ahissar

et al., 1992; Fuster, 1997; Tsumoto, 1992). If connected neurons fire together,

their mutual influence on each other becomes stronger.

Figure 2.5 illustrates the change of synaptic effectiveness caused by coac-

tivation of neurons (Kandel, 1991). A neuron in the hippocampus, a phylo-

genetically old structure belonging to the cortex, was investigated. If a singleinput fiber to the neuron was activated for some time, this did not change

connection strength. However, when the single input fiber was repeatedlyactivated together with several others so that the target neuron was caused

to fire together with the input, later inputs through the participating synapses

had a much stronger effect. This indicates that these synaptic connections

Figure 2.5. Illustration of long-term potentiation. (a) If a single input fiber (uppermost line)to a neuron was activated for some time, this did not change synaptic connection strength.(b) When the single input fiber was activated repeatedly with several others so that thetarget neuron was caused to fire together with the input, later input through the participatingsynapses was more effective. (c) Activation of the strong input alone selectively increased theinfluence of the stimulated neurons, but not that of others. Reprinted with permission fromKandel, E. R. (1991). Cellular mechanisms of learning and the biological basis of individuality.In E. R. Kandel, J. H. Schwartz, & T. M. Jessell (Eds.), Principles of neural science (3 ed.,pp. 1009–31). New York: Elsevier.

Page 37: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 37/331

20 Neuronal Structure and Function

became stronger. The increase of synaptic effectiveness was specific. Ac-

tivation of the strong input alone did selectively increase the influence of 

the stimulated neurons, but not that of others. Thus, the strengthening of synaptic links must have been a consequence of coactivation of the pre-

and postsynaptic neurons and also exhibited some specificity. The Hebbianspeculation receives support from these and similar data.

That neurons firing together wire together is, however, not the whole story.

Whereas neurons become associated when being activated repeatedly at the

same time, their antiphasic activation results in weakening of their influence

on each other (Tsumoto, 1992). Thus, it appears that it is the correlation of 

neuronal firing of connected cells, or a related measure, which is, so to speak,translated into their connection strength. The correlation c of the firing F of 

two neurons α and β can be calculated as a function of their probability to

fire together in the same small time window, op[F (α), F (β)], divided by their

individual likelihood to become active, p[F (α)] and p[F (β)]. For example,

the following formula can be used for calculating the correlation:

c = p[F (α), F (β)]

 p[F (α)] ∗ p[F (β)]

If the strength or weight of a synaptic connection between neurons is deter-

mined by the correlation coefficient, this implies (i) that neurons that fire

together frequently (high correlation) become more strongly associated, and

(ii) that neurons firing independently of each other (low correlation) will be

less strongly connected. Connection strength between pyramidal neurons inthe cortex is likely strongly influenced by the correlation of their neuronal

firing or a closely related measure.Modification of the strength of synaptic connections can be related to

biochemical and even structural changes in the neurons (Kandel, 1991), for

example, to the growth and modification of dendritic spines (Braitenberg &

Schuz, 1998). It should also be mentioned that coactivated of neurons maylead to formation of new synapses and spines (Engert & Bonhoeffer, 1999).

It should therefore be clear that, although the rule proposed by Hebb is im-

portant, it is not the only factor determining functional connectivity in the

cortex.

2.3 Principles and Implications

Thehumancerebralcortexisanetworkofmorethan10billionneurons.Each

neuron represents an information processor whose output is a function of 

the input from the many other neurons with whom it is interwoven. Based

Page 38: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 38/331

2.3 Principles and Implications 21

on the findings discussed so far, the following principles appear to reflect

universal neuroanatomical and neurophysiological properties of the cortex:

(I) Afferent and efferent projections are ordered.

(a) They reach, or insert from, primary areas (Figure 2.3).

(b) Sensory and motor projections are organized topographically

(Figure 2.4).(II) Intracortical connections permit mixing of afferent and efferent infor-

mation.

(a) Adjacent neurons are heavily connected and form local clusters.

(b) Primary areas tend not to be linked directly, but through relayareas.

(c) Adjacent areas are connected with high probability (>70%).(d) There is a lower but still good chance for connections between

areas farther apart (15–30 percent).

(e) Homotopic areas of the two hemispheres tend to be connected.

(f) Connections between areas tend to be reciprocal.

(III) Synaptic connections between neurons are modified depending on

their activity.(a) Neurons that fire together strengthen their mutual connections.

(b) Neurons that fire independently of each other weaken their con-

nections.

Clearly, this list of structural and functions features of the cortex is not

complete and more detail could easily be added. However, these basic facts

already make it possible to address more fundamental questions about the

function of the cortex.The cortex is supplied with information ordered according to modality

and, within each modality, ordered topographically. The cortex itself now

provides multiple indirect links between sensory- and action-related neu-

rons. These links are characterized by great divergence: Each neuron reaches

thousands of others and each area reaches tens of its sister areas. They arealso characterized by great convergence; that is, each neuron receives in-

put from multiple other neurons, and each area receives input from several

other ones. It has been argued by neuroanatomists (Braitenberg, 1978b;Braitenberg & Schuz, 1998) and neurocomputational modelers (Palm, 1982,

1993a) that this architecture is ideal for mixing and merging information

immanent to sensory- and action-related activity patterns. Because the map-ping between primary areas is indirect, through relay neurons and areas, it is

possible to store complex relationships between input and output patterns.

The correlation learning principle (III) implies that frequently cooccur-

ring patterns of activity can be stored by way of strengthening of synaptic

Page 39: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 39/331

22 Neuronal Structure and Function

links between the participating neurons. Such synaptic strengthening can

certainly take place between closely adjacent cortical neurons, but there

is reason to assume that such associative learning is also possible betweenneurons in distant areas. Even coactivation of neurons in different primary

cortices causing additional neuronal activity of neurons in their connectingrelay areas may be stored by way of increasing synaptic weights between

the simultaneously active units. This suggests that webs of neurons can form

that are distributed over various cortical areas.

In summary, it appears that the cortex can serve the function of merg-

ing multimodal information. This merging of multimodal information is not

done by direct links between primary areas, but necessitates intermediateneuronal steps. The intervening neuronsbetween sensoryand motor neurons

in the cortex allow for complex mappings of information patterns between

modalities. Such mappings can be stored by correlation learning.

2.4 Functional Webs in the Cortex 

The cortex is a network of neurons characterized by ordered input and

output connections in modality-specific areas, by heavy information mixingthrough short- and long-distance connections, and by correlation learning.

Such a device can serve the function of linking neurons responding to spe-cific features of input patterns and neurons controlling aspects of the motor

output. Because different primary areas are not linked directly, additional

neurons in nonprimary areas are necessary to bridge between the ordered

input and output patterns. The cortical connection structure, characterized

by a high connection probability between adjacent areas and more selective

long-distance links, allows for the formation of functionally coupled but dis-tributed webs of neurons reaching from the primary areas into higher-order

cortices. Development of these networks, called functional webs, would be

driven by sensorimotor or sensory–sensory coactivation, and would be de-

termined by the available cortical projections indirectly connecting the coac-

tivated neurons in primary areas to each other.

2.4.1 Why Numerous Neurons Should Cooperate

It was pointed out by Donald Hebb (1949), and this may be his most impor-

tant contribution to the understanding of the brain, that synchronously acti-vated neurons should link into cell assemblies, and that cell assemblies under-

lie all higher cognitive processes. Hebb’s proposal diverged radically from

earlier neuroscientific approaches to information processing in the brain,

because he postulated that higher brain processes are realized as functional

units above the level of the neuron. Earlier proposals had stated that either

Page 40: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 40/331

2.4 Functional Webs in the Cortex 23

individual neurons (Barlow, 1972) or mass activity and interference patterns

in the entire cortex (Lashley, 1950) are the basis of cognition. Hebb’s view

may appear as a compromise between these views (Milner, 1996).While Lashley’s proposal can be ruled out by considering the specific neu-ropsychological deficits caused by focal brain lesions (Shallice, 1998), one

may ask why large neuron ensembles should become involved in cognitive

processing if single neurons are already capable of performing the relevant

computations. A tentative answer is that individual neurons are too noisy

and unreliable computational devices so that it is advantageous to use sets of neurons working together in functional units to achieve more reliable infor-

mation processing. If the signal-to-noise ratio of individual neurons is low,one can obtain a better signal by simultaneously averaging a larger number

of neurons with similar functional characteristics, so that uncorrelated noise

is cancelled (Zohary, 1992). (This does not rule out the possibility that, apart

from their shared function, individual neurons in the ensemble can haveadditional specific functions.) It would therefore make good sense if there

were functional units in the cortex that are larger than the neuron but much

smaller than the neuronal populations in the macroscopic cortical gyri and

sulci.A further argument in favor of functional webs composed of numerous

neurons comes from an estimate of the number of neurons necessary for car-ryingoutthetasksthecortexseemstobeprimarilyengagedin.Asmentioned

earlier, the cortex includes>10 billion neurons. The number of to-be-stored

items can be estimated on the basis of the units that need to be stored. To

speak a language well, one needs a vocabulary of fewer than 100,000 words

or meaningful language units, called morphemes, and a limited set of rules

governing their serial order (Pinker, 1994). Given similar numbers of distinctrepresentations also developed for other cognitive domains, the number of 

to-be-organized engrams may be in the order of a few 100,000. If this esti-

mate is correct and each engram is represented by one neuron, one million

individual neurons might be sufficient for representing the various percepts

and motor programs cognitive processes operate on. This raises the question

why there are 100,000 to 1,000,000 times as many neurons as would be nec-essary, as these considerations would suggest. A possible answer is that the

cortex includes so many neurons, because individual engrams are realizedas populations of neurons of 105 to 106 neurons.

2.4.2 The Need for Connecting Neurons in Distant Cortical Areas

Functional units in the cortex above the level of single cells are the already

mentioned local clusters of neurons beneath approximately 0.1–0.5 mm2

of cortical surface that, in various sensory areas, respond to similar stimuli

Page 41: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 41/331

24 Neuronal Structure and Function

(Hubel, 1995). From the perspective of cognitive science, however, these lo-

cal neuron clusters per se cannot be the substrate of the linkage between

different features of an object. The different features of an object maycharacterize input from different modalities, as, for example, the shape,smell, purr, and smooth fur of a cat. The binding of these features into one

coherent representation could be instantiated by pathways linking the sen-

sory information from different modalities to the same “central” neuron(s).

These critical cells should then be housed in areas where inputs from many

sensory fields converge. It is, however, not necessary to assume a cardinal 

cell , or central convergence area.

As noted, Barlow (1972) has postulated such “grandmother” or “cardinalcells” and Damasio (1989) has proposed that categories of knowledge are

processed not by single neurons, but in single “convergence zones” in the

cortex. Although these proposals are certainly helpful for understanding

aspects of brain function, they do not exhaust the range of possible brainmechanisms. The neuroanatomical connection pattern of thecortexindicates

that links between primary cortices are provided through more than one

route, involving several nonprimary areas. There is, therefore, no need to

assume single specialized areas or neurons for binding of defined entities.Therefore, a distributed web of neurons that processes a certain type of 

information may include a large number of specialized cells distributed overseveral cortical areas.

2.5 Defining Functional Webs

A web of neuronal links strongly connecting all neurons involved in the

specific processes triggered by an object in the input may become the corticalrepresentation of this object. In this case, binding of object features would

be established by mutual links within a distributed functional web, that is,

between neurons in widespread areas including the primary areas. Each

neuron member of the web would therefore contribute to holding the web

together thereby playing an essential role in its functioning. The cat concept

would be realized as a large set of neurons distributed over a small set of cortical areas. All of these areas would serve as binding sites. A functional

web will be assumed to be a set of neurons

(i) That are strongly connected to each other

(ii) That are distributed over a specific set of cortical areas

(iii) That work together as a functional unit 

(iv) Whose major parts are functionally dependent on each other so that

each of them is necessary for the optimal functioning of the web.

Page 42: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 42/331

2.5 Defining Functional Webs 25

The term functional web is usually preferred to the term cell assembly in

the context of this book. The established term cell assembly is sometimes

avoided because it has been used in so many different ways by different re-searchers that misunderstandings appear unavoidable (see also Chapter 8).

In contrast, the use of the new term functional web has now been clarifiedso that misunderstandings should be unlikely.

To speak about a selection of neurons without implications such as

(1) to (4), the expression neuronal group is sometimes used. The mem-

bers of a neuronal group would therefore not need to be connected to each

other. Established terms such as cell assembly, neural assembly, or neuronal 

ensemble are used without the specific implication that (4) is necessarily true.In addition, it is argued that functional webs can exhibit different types of ac-

tivity states, ignition and reverberation (Section 2.8). This is not a necessary

assumption but is, as is argued, supported by empirical data. In Chapter 10,

more specific types of functional webs are proposed that may be relevant

for grammar processing in the brain.

Whenever one tries to define exactly a given cell assembly or functionalweb within a simulated or real neuronal network, this turns out to be difficult.

In simulations using artificial associative memories (Palm & Sommer, 1995),one immediately encounters the problem of determining which neurons be-

long to the assembly and which do not. What one finds are neurons that are

connected to many of their fellows by maximal synaptic strength. For these,

it is out of question that they are assembly members. However, there are alsoothers whose connections to the rest of the assembly is slightly weaker and

whose inclusion in the assembly is therefore uncertain. Correspondingly, in

network simulations and also in neurophysiological observations, some neu-

rons almost always become active together, but others may only be recruitedtogether with them with lower probability, depending, for example, on the

neuronal sets activated in the past (Milner, 1957). In other words, there areneurons exhibiting high correlation among each other and others whose cor-

relation with these “core neurons” is smaller. To decide whether a neuron is

part of an assembly or web, critical values for synaptic connectedness or cor-

relation coefficients must be introduced. Definition of such critical values is

always arbitrary. For some purposes, it helps to distinguish the kernel or core

of an assembly from its halo or periphery (Braitenberg, 1978a), but in net-work simulations, necessarily arbitrary boundaries must also be introduced

for defining these assembly parts.

Cell assemblies and functional webs are necessarily fuzzy concepts. This

does not constitute a principled problem. It is essential to see that fuzziness is

immanent to the assembly concept and that this is problematic in exactly the

same way as it is difficult to determine the boundaries of the sun or the milky

Page 43: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 43/331

26 Neuronal Structure and Function

way. The fuzziness of the boundaries of the respective concepts should not

obscure the fact that there are stars, galaxies, and perhaps functional webs

in the brain that may represent these concepts and words.Perhaps most importantly in the context of a cell assembly approach to

language, the meaning of words in ordinary language use also is characterizedby the lack of sharp boundaries (see Chapter 5; Wittgenstein, 1953). If our

conceptual system is fuzzy, it is probably best to model it using fuzzy mecha-

nisms and to speak about it using words with fuzzy meaning. The anatomical

fuzziness of the boundaries of the individual functional web should not ob-

scure the fact that, functionally, each web is conceptualized as a coherent

unit, a discrete entity.

2.6 Evidence for Functional Webs

Which critical predictions are implied by the idea of distributed functional

webs? If the cat concept is neuronally represented as a distributed web of 

neurons that form a functional unit, this has two important implications:

(1) A significant portion of the web’s neurons are active whenever the catconcept is being processed.

(2) The function of the web depends on the intactness of its member

neurons.

If neurons in the functional web are strongly linked, they should showsimilar response properties in neurophysiological experiments. If the neu-

rons of the functional web are necessary for the optimal processing of the

represented entity, lesion of a significant portion of the network neuronsmust impair the processing of this entity. This should be largely indepen-

dent of where in the network the lesion occurs. Therefore, if the functional

web is distributed over distant cortical areas, for instance, certain frontaland temporal areas, neurons in both areas should (i) share specific response

features and (ii) show these response features only if the respective other

area is intact.

These predictions have been examined in macaquemonkeys using a mem-

ory paradigm in which the animal must keep in mind the shape or color of astimulus and perform a concordant matching response after a delay of sev-eral seconds (delayed matching to sample task). Throughout the memory

period, in which the animal must keep in mind, for example, that the stimu-

lus shown was red, neurons fire at an enhanced level. Their firing is specific in

the sense that they do not respond, or respond less, when a stimulus of a dif-

ferent color is shown. Neurons with this stimulus-specific response pattern

Page 44: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 44/331

Page 45: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 45/331

28 Neuronal Structure and Function

or, more precisely, engram-related neurons in one area led to functional

impairment of the neurons in the respective other area (Fuster, 1997). To-

gether, these data provide evidence that neurons in both temporal andfrontal areas (a) showed the same specific response features and (b) showed

these response features if and only if the respective other area was intact(Fuster, 1997).

These results obtained in memory experiments with macaque monkeys

are reminiscent of well-known facts from more than 100 years of neurolog-

ical investigation into acquired language disorders, aphasias (Broca, 1861;

Lichtheim, 1885; Wernicke, 1874). These classical studies of aphasia showed

that prefrontal and temporal areas are most crucial for language processing.They also showed that lesions in either area can lead to aphasia, which in the

majority of cases include deficits in both language production (Lichtheim,

1885) and perception (De Renzi & Vignolo, 1962). Concordant with recent

animal studies investigating the consequences of local cooling of prefrontal

and temporal areas, this suggests mutual functional dependence between

frontal and temporal areas (Pulvermuller & Preissl, 1991).

2.7 A View of Cortical Function

The reviewed facts from basic neuroscience make it likely that the cortex

includes distributed neuron ensembles involving neurons in cortical areas

in different lobes that show similar response properties and whose intact-

ness is necessary for defined cognitive operations. This further advocatesthe following view of cerebral cortical function. The cortex is an associa-

tive memory allowing for merging information from different modalities.

Merging of information is driven by correlation of spatiotemporal patternsof neuronal activity carrying information about sensory perceptions and ac-

tions. These correlated activity patterns occur in action-related frontal and

sensory-related posterior areas of the cortex. The participating neurons arebeing bound into strongly connected webs of neurons, functional units that

represent cognitive entities with sensory and action aspects (words, concepts

engrams in general).

This view on cortical function leads to further questions, as follows:

(1) What  are the functional dynamics of these distributed neuronalrepresentations?

(2) Where exactly are they localized, or, formulating the question in a

slightly more adequate manner: Over which cortical areas are these

functional webs distributed, i.e., what is their cortical topography?

(3) How can the internal wiring of the functional webs be specified?

Page 46: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 46/331

2.8 Temporal Dynamics in Functional Webs 29

(4) Is it sufficient to assume that the formation of these networks is driven

by associative learning principles, or do genetically determined factors

play a role as well?The “What” question (1) about functional dynamics is answered tentativelyin the last section of this chapter and in Chapters 10 and 12. The “Where”

question (2) is addressed in detail in Chapters 4 and 6. The “How” question

(3) receives a tentative answer in Chapter 8 on serial order, and the last

question (4) about the possible role of genetically determined information

is likely impossible to answer, but an attempt is made in Chapters 4 and 14.

2.8 Temporal Dynamics in Functional Webs:Ignition and Reverberation

On theoretical grounds, one can predict that a strongly connected neuron

population produces overshooting activity if strong stimulation reaches it

from the outside. This would mean that, although only a fraction of the

neuron members of a web are stimulated, all members of the web, or at

least most of them, would nevertheless become active, and activity mighteven spread to other networks as well. Restricting considerations here to

the activity dynamics within the functional web, stimulation of a fraction of 

its neurons can lead to a full activation of the entire population. This process

has been called ignition (Braitenberg, 1978a). Neural network models using

associative memory models leave no doubt that such a process takes place ina sufficiently strongly connected and sufficiently strongly stimulated network

of neurons (Palm, 1981, 1982).

If activation of a substantial fraction of the neurons of a functional webleads to its ignition, one may consider this an undesired consequence of the

architecture of network memories. However, the assembly internal spread

of activity may as well be considered as positive because the activation of the web, so to speak, completes itself as a result of the strong web-internal

links. If the web of neurons is considered a memory representation of an

object and each neuron to represent one particular feature of this object

memory, the full ignition would be the neuronal correlate of the activation

of the stored object representation. Such full activation of the object memorycould occur if only a fraction of the features of the object are present in theactual input. This could be a psychologically important process, an organic

correlate of the completion of a gestalt in the perception process. Gestalt

completion can be modeled in associative memories (Willshaw, Buneman,

& Longuet-Higgins, 1969), and there is reason to believe that a similar pro-

cess is relevant in the real brain.

Page 47: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 47/331

30 Neuronal Structure and Function

The dynamics of a functional web would therefore be characterized by full

ignition of the web after sufficiently strong stimulation by sensory input, or,

as an alternative, after stimulation by cortical neurons outside the functionalweb. This latter cortex-internal activation of a web can be considered the

organic basis of being reminded of an object even though it is absent in theenvironment.

The ignition process likely takes place within a short period of time after

the functional web has been sufficiently stimulated. An exact definition of 

the interval is probably not sufficiently motivated by empirical data, but an

educated guess might be that the ignition occurs within 200 ms after sufficient

information is present in the input. The issue of temporal dynamics of mem-ory representations is further discussed in the context of neurophysiological

activity accompanying word processing (see Chapter 4).

When an ignition has taken place, refractory periods of neurons and other

fatigue effects may again reduce the activity level of the functional unit. It is

also possible that a regulation process designed to keep the cortical level of 

activity within certain bounds (see Chapter 5) becomes active and reducesactivity by global inhibition or removal of background activity (disfacilita-

tion). In this case, however, activity will not necessarily completely die outin the ensemble of neurons that just ignited. The strong within-assembly

connections may still allow activity to be retained in the neuron set. This is a

putative neurobiological basis of short-term or active memory (Fuster, 1997,

1998a, 1998b, 2000). The distributed cortical functional web itself wouldtherefore be the organic side of a long-term (or passive) memory trace, and

thesustainedactivityofthesamewebwouldrealizetheshort-term(or active)

memory.

Researchers who have found memory-specific neurons in the neocortexhave investigated the dynamics of these cells extensively while the experi-

mental animal must keep a certain piece of information in mind. The animalsaw a stimulus of a certain color and had to push a button of the same color

after a memory period of 20 seconds. Throughout the entire memory pe-

riod, neurons were found that stayed at an enhanced activity level. Some of 

these neurons were stimulus specific, therefore responding strongly only if 

the to be remembered stimulus exhibited a particular feature. The neuron

did, for example, respond strongly to a yellow stimulus, but did not show anyenhanced activity, or at least responded significantly less, when the stimulus

had a different color, for instance, green or blue (Fig. 2.6). The explanation

is that ensembles of neurons were activated by the input and stayed active

during the memory period. The activity dynamics of the individual mem-

ory cell would be assumed to reflect the activity dynamics of the memorynetwork. The enhanced activity level of the neuron and web would be the

Page 48: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 48/331

GREENSAMPLETRIALS

REDSAMPLETRIALS

YELLOW

SAMPLETRIALS

BLUESAMPLETRIALS

    S    P  I    K    E    S  /    S    E     C

25

20

15

10 5 0

BASELINE

S

TIME(SE

C)

DELAY

M

POST-CHOICE

0

5

10

15

       F       i     g   

     u      r     e 

       2  . 

       6   . 

     R    e     s    u 

     l    t      o      f     a 

    n    e    x

    p      e     r     i    m    e     n    t 

      i    n   w

     h     i    c      h    a     m    o     n

     k    e    y  

      h    a 

     d     t     o 

     k    e     e     p  

     i    n    m

     i    n     d      i     f

     a     g      r    e     e     n

 ,      r    e 

     d  ,     y  

    e      l     l    o 

   w ,      o 

    r      b      l

   u     e     s     a     m    p  

     l    e     s     t      i    m   u 

     l   u     s    w

    a     s 

     s      h    o    w

    n     d 

   u     r     i    n    g  

    t      h    e     s     t      i    m   u 

     l    a     t      i    o     n

     i    n    t     e     r   v    a 

     l      S 

 .       D

   u     r     i    n    g      t      h    e     m    e     m    o     r   y      p      e     r     i    o 

     d      (      “     d     e 

     l    a    y  

     ”     )  , 

     t      h    e     n    e    u 

    r    o     n

     f    r    o     m

   w     h     i    c      h    r    e     c     o     r     d 

     i    n    g      s 

    w    e     r    e     t     a 

     k    e     n    r    e     s     p      o     n

     d     e 

     d     m    a    x

     i    m    a      l     l   y      a 

     f    t     e     r

     a 

   y      e 

     l     l    o    w

    s     a     m    p  

     l    e     s     t      i    m   u      l   u 

    s  .       H

     i    g       h    a     c     t      i   v

     i    t    y      o 

     f     t      h

    e     n    e    u     r

    o     n    a     p      p      e     a     r    s     t     o 

     b     e     a     n

     i    n     d      i    c     a     t     o     r

     o      f

     a     s     p      e     c 

     i     fi    c     e     n    g      r    a     m

     i    n    m

    e     m    o     r   y  

 .       D

   u     r     i    n    g      t      h    e 

     d     e 

     l    a    y  

 ,      t      h

    e     n    e    u 

    r    o     n

     l    o     s     e     s 

     a     c     t      i   v

     i    t    y       a 

     l    m    o     s     t      e 

   x    p      o     n    e     n    t      i    a 

     l     l   y   .       R

    e     p      r     i    n    t     e 

     d    w

     i    t      h

    p      e     r    m

     i    s     s 

     i    o     n

     f    r    o     m

     F   u 

    s     t     e     r , 

      J  .       M

 .       (      1

     9     9     5      )  .     M   e    m   o    r   y  

     i   n     t      h

   e     c    e 

   r   e      b 

   r   a      l

    c    o    r    t    e    x

 .      A

   n    e 

   m   p  

    i   r    i   c    a      l

    a    p     p     r   o    a    c 

     h

    t    o     n

   e    u    r   a 

     l    n

   e     t    w   o    r     k   s 

     i   n     t      h

   e       h

   u    m   a    n

    a    n

     d     n

   o    n

     h   u    m   a    n    p  

   r    i   m   a 

    t    e  .       C 

    a     m

     b     r     i     d     g      e 

 ,       M

     A    : 

      M     I     T     P    r    e     s     s 

 . 

31

Page 49: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 49/331

32 Neuronal Structure and Function

result of the reverberation of activity within the functional web. The stimulus

specificity of the neurons suggests memory networks specific to objects or

object features.This research revealed sometimes complex activity dynamics. As men-

tioned earlier, neurons usually did not stay at a constant enhanced level of activity; instead, gradual activity changes were found. Many neurons showed

strong activation after stimulus presentation followed by an exponential de-

cline of activity with time throughout the memory period. Figure 2.6 illus-

trates this. The activity during the memory period – in action potentials per

second – is displayed as a function of time when stimuli of different color

had to be memorized. These data can be interpreted as support for the ideathat after the ignition of a cortical neuron ensemble corresponding to the

stimulus, this network stays at an enhanced activity level for several tens

of seconds. It has been proposed that activation of a distributed cortical

neuron ensemble is reflected by coherent oscillatory activity of the partic-

ipating neurons. This claim receives support from both neurophysiological

studies in animals (Kreiter, 2002; Singer et al., 1997) and EEG and MEGstudies in humans (Pulvermuller, Birbaumer, Lutzenberger, & Mohr, 1997;

Tallon-Baudry & Bertrand, 1999).In summary, neurophysiological data and neurocomputational investiga-

tions suggest complex activity dynamics of functional webs. An initial brief 

ignition of the assembly appears to be followed by a reverberatory state in

which the assembly retains activity, although the level of activity may fall off exponentially with time. This memory interval of reverberatory activity can

last for tens of seconds.

Page 50: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 50/331

CHAPTER THREE

From Classic Aphasia Researchto Modern Neuroimaging

In this chapter, an introduction to facts known from neurological, neuropsy-

chological, and neuroimaging research on language is given. How the ob-

tained results can be explained and how they fit together is discussed.

3.1 Aphasiology

The scientific study of language proliferated in the second half of the nine-

teenth century. Apart from linguists and psychologists, neurologists focused

on language in the context of the then new findings about “affections of 

speech from disease of the brain” (Jackson, 1878). In adults who had been

fully able to speak and understand their native language, a stroke, tumor,trauma, or encephalitis was sometimes found to severely and specifically re-

duce their language abilities. Such language disturbances were called

aphasias. There was, and still is some discussion as to whether there aresubtypes of aphasia, and a good deal of the research on aphasia was dedi-

cated to developing new classification schemes and arguing why one scheme

should be a better reflection of what appeared to be the truth than another.The research effort resulted in numerous classification schemes (Caplan,

1987) and also in what appeared to be an extreme position expressed by the

claim that there is only one type of aphasia (Marie, 1906).

This last view can be based on the common features of all – or at least

the large majority of – aphasias. All aphasics have difficulty speaking, al-though their ability to move their articulators – lips, tongue, pharynx, larynx,and other muscles in the mouth-nose region – may be well preserved. When

confronted with spoken and written language, all aphasics exhibit some dif-

ficulty. An easy way of finding out whether an individual suffering from

brain disease has an aphasic disturbance is to present him or her with a few

simple commands and assess whether the patient is still able to carry them

33

Page 51: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 51/331

34 From Classic Aphasia Research to Modern Neuroimaging

out. A test called the Token Test (De Renzi & Vignolo, 1962) includes a list

of commands to manipulate colored tokens of different shape and size. If 

a patient has an aphasia, he or she usually shows difficulty understandingand following commands such as, “Please touch the little yellow circle,” or,“Please take all the squares except for the red one.” In addition to their

comprehension problems, all aphasics have difficulty producing spoken and

written language output. It has been reported that there are exceptional

patients who would show a deficit only in oral language production, or a

deficit only in comprehending certain aspects of spoken language (Kolk,van Grunsven, & Keyser, 1985). Apart from these few exceptions, however,

aphasias appear to be multimodal ; that is, they include deficits in producingand comprehending spoken language and additional deficits in reading and

writing as well. The multimodal character of aphasia can be used to argue

in favor of the view that there is only one type of aphasia, aspects of which

happen to be more or less pronounced in different cases.However, if the focus is on the diversity of the actual problems patients

with acquired language deficits exhibit, it makes sense to distinguish dif-

ferent syndromes of aphasia. The most obvious differentiation is between

aphasics whose primary difficulty is producing language and aphasics whoshow a most massive comprehension deficit. Here, a distinction betweenmotor aphasia and sensory aphasia has been proposed. Today, the termsBroca aphasia and Wernicke aphasia – referring to these variants of the dis-

ease, respectively, and to the two neurologists who first described them – are

established. Apart from these two aphasic syndromes, a severe form of apha-

sia, global or total aphasia, and a slight disturbance characterized mainly by

word-finding problems, amnesic aphasia or anomia, are sometimes distin-

guished. More fine-grained syndromes of aphasia have been proposed onthe basis of theoretical approaches of language processing in the brain.

In the second half of the nineteenth century, neurologists proposed con-

nectionist theories to summarize and model the effect of brain lesions on

cognitive functions. The main idea underlying these models was that local

processing systems, or centers, specialize in particular cognitive processes.

These were believed to be autonomous processors realized in large portionsof gray matter, such as the superior temporal gyrus or the inferior frontal

gyrus. The centers were thought to be linked through pathways allowing in-formation exchange between them. The pathways in the models had their

analog in fiber bundles in the white matter of the brain.

The most famous classical neurological model of language goes back to

Broca (1861) and Wernicke (1874) and has been proposed by Lichtheim(1885). This model is consistent with basic features of aphasia. The main

proposal of Lichtheim’s model was that there are two centers for language

Page 52: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 52/331

3.1 Aphasiology 35

Table 3.1. Nomenclature of some areas relevant for language processing.Names of areas are indicated and related to the number code proposed

by Brodmann. (See Fig. 2.3 for the corresponding loci in Brodmann’s areamap.)

Brodmann

Name number Latin name Location

Broca’s area 44 Gyrus opercularis Inferior frontal lobe

45 Gyrus triangularis Inferior frontal lobe

Wernicke’s area 22 Gyrus temporalis Superior temporal

superior lobe

39 Gyrus supramarginalis Inferior parietal lobe

40 Gyrus angularis Temporo–parieto–

occipital junction

processing: one involved in speech production and the other in languagecomprehension. The existence of two separate anatomical centers relevant

for language, the areas of Broca and Wernicke, has meanwhile been con-

firmed by many clinical observations. These areas are located in the left

hemisphere in most individuals, Broca’s area in the inferior frontal gyrus,

Brodmann areas 44 and 45; and Wernicke’s area in the superior temporal

lobe, Brodmann area 22. Both are located close to the Sylvian fissure, thatis, in the perisylvian region (Table 3.1).

Figure 3.1 presents Lichtheim’s model of language representation and

processing in the brain. He distinguished three centers, two of which are

devoted exclusively to the processing of aspects of language. The first cen-

ter, the motor language center M, was supposed to store what Lichtheim

Figure 3.1. Lichtheim’s model of language in thebrain. Language was proposed to be realized inthree centers involved specifically in representingand processing sound patterns of words (Wortk-langbilder ; acoustic language center A), articu-latory patterns of words (Wortbewegungsbilder ;motor language center M), and concepts (Begriffe ;concept center B). Adopted from Lichtheim, L.(1885). ¨ Uber Aphasie. Deutsches Archiv f   ur Klin-ische Medicin, 36, 204–68.

Page 53: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 53/331

36 From Classic Aphasia Research to Modern Neuroimaging

called Wortbewegungsbilder , or representations of the articulatory move-

ments performed to pronounce a word. The second center, the auditory

language center A, was thought to store what Lichtheim calledWortklang-

bilder , or images of the sound sequences of words. The third center, which

was hypothesized to play a major role in other mental activities as well, wascalled the Begriffszentrum, the center storing concepts, B. Any two of these

three centers were proposed to be connected by pathways allowing for con-

verting a sound sequence directly into a sequence of muscle movements or

a concept into a word and back. The genuine language centers A and M

correspond to the areas of Broca and Wernicke.

Language processes were modeled by activity flowing through the connec-tionist architecture. For example, word comprehension would be modeled

as the activation of A by acoustic input and the subsequent activation of B

through its connection to A. The cortical process of language production,

however, was mimicked by activity in B, exciting, through the B–M pathway,

the motor center M, and finally causing a motor output. The process of 

thoughtless repetition of a word was believed to be possible without activat-ing B, based on the direct connection between the acoustic and the motor

centers. Aphasias were now modeled as a lesion to one or more centers orpathways. Broca and Wernicke aphasias, for example, were realized in the

model as lesions of the respective centers M and A.

The nice thing about Lichtheim’s scheme was that it suggested additional

syndromes of aphasia that could be confirmed empirically. Examples areconduction aphasia, modeled as lesion of the connection from A to M,

a syndrome primarily characterized by difficulty in repeating word se-

quences, and a mirror-image syndrome called mixed transcortical aphasia,

in which repetition is still relatively intact whereas all other language abili-ties including the ability to understand what is being repeated are severely

impaired.Although the existence of two centers whose lesion in many cases leads to

severe language disturbance could be confirmed, it is less clear that these cen-

ters have the specific function Lichtheim and Wernicke attributed to them.

As Lichtheim recognized in 1885, it is incorrect to assume that what has

been labeled the “center for language comprehension” is relevant only for

this single language function. As emphasized earlier, patients with Wernickeaphasia have difficulty speaking, even if their lesion is restricted to the supe-

rior temporal lobe (Pulvermuller, Mohr, Sedat, Hadler, & Rayman, 1996).

They use words in inappropriate contexts; produce words incorrectly, with

incorrect language sounds in them or language sounds omitted; or even pro-

duce an incomprehensive jargon of mixed-up language sounds. These deficits

are typical for subtypes of sensory or Wernicke aphasia, and cannot be easily

Page 54: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 54/331

3.1 Aphasiology 37

explained by assuming a selective lesion to a center devoted to language

comprehension. Furthermore, the neurologically caused speech errors fre-

quently made by Wernicke aphasics differ qualitatively from speech errorsmade by Broca aphasics, who tend to speak very slowly because they cannot

produce words they intend to use, rather than replacing them with wrongones. A lesion in an “auditory language center” cannot account for such

differential output. A similar argument can be made for the comprehension

side, because in Broca aphasia, specific comprehension deficits are partic-

ularly apparent when patients are confronted with certain sentence types,

including, for example, passive sentences (Caramazza & Zurif, 1976). Thus, it

would be incorrect to postulate a cortical center specifically for language pro-duction and a second independent center processing auditory language inputexclusively. The two areas most crucial for language processing in the cortex,

the inferior frontal area of Broca and the superior temporal area of Wernicke,

appear to be functionally interdependent (see Chapter 2; Excursus E1).

Why are the language areas located where they actually are? It might be

tempting to propose an additional principle, adding to the ones proposedearlier in Chapter 2; a principle saying that the human cortex is constructed

such that language must be processed in these very areas. As an alternative,however, it is possible to explain the location of the language centers by

principles already known from neuroanatomy and neurophysiology.

It is clear that production of a language element, for example, a syllable

or word, corresponds to activation of neurons controlling movements of thearticulators. It is likewise evident that such articulations must always activate

neurons in the acoustical cortical system, unless the auditory system is dam-

aged or otherwise dysfunctional. Thus, when uttering a language element,

there must be correlated neuronal activity in the perisylvian motor cortexand auditory cortex in the superior temporal lobe. Because between-area

connections preferably connect adjacent areas, this neuronal activity canspread to adjacent fields in the superior temporal and inferior frontal lobes.

These areas are connected by long-distance connections. Thus, a sufficiently

strong correlated activity pattern in primary motor and auditory areas gives

rise to the activation of a specific neuron population spread out over these

inferior frontal and superior temporal areas, including neurons in the lan-

guage areas of Broca and Wernicke. The principles outlined in Chapter 2 aresufficient for explaining why the major language areas are where they are.

The outlined view of the development of sensory-motor links in language

acquisition has been established for some time (Braitenberg, 1980; Fry, 1966).

It is sometimes called the babbling hypothesis, because during the infant’s

repeated articulation of syllables within the first year of life, this type of asso-

ciative learning may occur for the first time. Strong associative links between

Page 55: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 55/331

38 From Classic Aphasia Research to Modern Neuroimaging

the two principal language areas predict that these areas are not functionally

independent, but that they become active together and cooperate when they

generate, or respond to, language sounds, words, or sentences. This cooper-ation postulate contrasts with the view immanent to the Lichtheim scheme

that the frontal and temporal language areas are autonomous centers. Strongassociative links between the areas of Broca and Wernicke may be the basis

of the multimodal character of most aphasias, a feature Lichtheim’s scheme

cannot account for. This issue was touched upon earlier and will be discussed

further in detail in Excursus 1.

One may object to this view by mentioning that in users of sign language,

large lesions in the perisylvian areas also lead to aphasia (Poizner, Bellugi,& Klima, 1990). Should not the frontal language center for sign languages

be located superior or dorsal to that of spoken language because signing is

done primarily by moving one’s hands rather than the face and articulators?

Recall that the somatotopic organization of the primary motor cortex im-

plies a more dorsal position at least for the neurons directly controlling arm

movements (see Fig. 4.4). However, because many of the lesions seen in deaf signers with aphasia were rather large and affected both the perisylvian ar-

eas and the more dorsal areas controlling hand movements, they cannot helpdecide the issue. There are very few exceptions – for example, the case of a

patient called Karen L., who suffered from a relatively small lesion dorsal

to Broca’s area but which extended into the supramarginal and angular gyri

(Poizner et al., 1990). Poizner and colleagues emphasize that a lesion of thiskind does not usually lead to persistent aphasia including comprehension

deficits, and attribute this patient’s aphasia to the posterior areas affected.

However, it is also possible to relate the deficit to the anterior part of the

lesion dorsal to Broca’s region. The data on aphasia in signers appear to beconsistent with an explanation of the cortical locus of the areas most rele-vant for language based on correlated neuronal activity and the available

neuroanatomical links.

Unfortunately, neuroimaging data addressing the issue of the cortical

areas involved in sign language processing are controversial at present.

Some colleagues report massive differences in the cortical organization of 

spoken and signed language, for example in the pattern of cortical lateral-

ity (Bavelier, Corina, & Neville, 1998; Neville et al., 1998; Newman et al.,2002). In contrast, others claim that the cortical locus of spoken language

and that of sign language are similar (Petitto et al., 2000). The interpretation

of brain correlates of spoken and sign languages is complicated by cortical

reorganization (Buonomano & Merzenich, 1998). If arms and hands rather

than the articulators are excessively used for communicating, general corticalreorganization processes may be triggered so that some of the articulators’

Page 56: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 56/331

3.2 Laterality of Language 39

cortical space is taken over by the extremities. This may explain why similar

areas are activated in the frontal lobe when signed and spoken languages

are processed (Petitto et al., 2000). Cortical reorganization may be an im-portant factor working against clearly separate cortical loci devoted to the

processing of spoken and signed languages.Whereas the language areas of Broca and Wernicke were for some time

believed to be the only areas crucial for language, neuropsychological work

carried out in the last quarter of the twentieth century proved that also

other areas are necessary for unimpaired language processing. In particular,

lesions in the frontal and temporal lobes, some of which spared the peri-

sylvian language areas, led to difficulty producing or understanding words.Many of these aphasic deficits were most pronounced for words from partic-

ular categories – nouns, verbs, or more fine-grained semantic subcategories

of words and concepts (Damasio & Tranel, 1993; Humphreys & Forde, 2001;

Warrington & McCarthy, 1983; Warrington & Shallice, 1984). For example,

frontal lesions appear to limit the ability to process verbs (Bak et al., 2001;

Daniele et al., 1994; Neininger & Pulvermuller, 2001), whereas inferior tem-poral lesions were found to most severely impair the processing of nouns

from certain categories (Bird et al., 2000; Damasio et al., 1996; Warrington& McCarthy, 1987). Clearly, these language-related deficits remain unex-

plained by the Lichtheim scheme.

3.2 Laterality of Language

Laterality of language is a well-known fact since the first scientific investi-

gation of language loss due to stroke (Broca, 1861), but the causes of this

laterality have not yet been revealed. The postulate that one hemisphere isdominant for language is based primarily on lesion studies. Only lesions in

the left hemisphere cause aphasias in most individuals. It was pointed out by

English neurologist Hughlings Jackson (1878) that if a lesion of a part of thebrain impairs specific functions, one can by no means conclude that these

functions are localized in the respective brain part. The lesioned area could

have a more general function, as the brain stem has in regulating arousal,

which is necessary for, but not specific to, a specific higher brain function such

as language. In this case, one would perhaps not want to localize languagein the brain part in question, although language impairments result from itslesion. Likewise, if lesions of a brain part lead to a clinically apparent deficit

regarding a given function, it is always possible that additional areas are also

relevant for this function, but that their lesion does not result in clinically

apparent dysfunction. Such deficits may be absent, for example, because

the clinical tests applied are not sensitive enough to reveal a fine-grained

Page 57: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 57/331

40 From Classic Aphasia Research to Modern Neuroimaging

dysfunction (Neininger & Pulvermuller, 2001), or because other areas had

meanwhile taken over the damaged area’s function (Dobel et al., 2001; Price

et al., 2001; Weiller et al., 1995). Therefore, Jackson’s early claims are correct:Loss of function F after lesion in area A does not prove that F is exclusivelyhoused in A, and the absence of an F-deficit as measured by clinical tests,

after lesion in A does not prove that A has no role in the respective function.

Therefore, lesion data proving language laterality do not argue against the

existence of additional sites in the nondominant hemisphere that are also

relevant for language processing.

Whereas lesions in certain left-hemispheric areas cause severe language

impairments and aphasias, right-hemispheric lesions usually lead to moresubtlelanguage-related difficulties affecting prosodic and pragmatic process-

ing (Joanette, Goulet, & Hannequin, 1990). In this sense, left-hemispheric

language dominance is almost always present in persons who are right

handed and also in most left-handed individuals (∼80 percent) (Bryden,

Hecaen, & DeAgostini, 1983; Goodglass & Quadfasel, 1954; Hecaen, DeAgostini, & Monzon-Montes, 1981). The remaining individuals are right

dominant, with a few showing no language dominance at all. Therefore, it

is obvious that in the large majority of individuals, language is lateralized tothe left hemisphere.

Language laterality was also reflected in brain physiology revealed by

modern neuroimaging techniques. Stronger brain responses in the left hemi-sphere in comparison with the right hemisphere were seen across various

tasks using visual or auditory stimuli (Petersen & Fiez, 1993). Because later-

alized activity was elicited already by single language sounds and syllables

(Naatanen et al., 1997; Shtyrov et al., 2000; Zatorre et al., 1992), one may

conclude that phonological processes, or more precisely, acoustic processesrelevant for the distinction of language sounds are crucial for language later-ality. In many of the neuroimaging studies mentioned, in particular in studies

using magnetoencephalography (MEG), multichannel electroencephalog-

raphy (EEG), or functional magnetic resonance imaging (fMRI), language

laterality was gradual, in the sense that there were activity signs in both

hemispheres and the dominant left hemisphere was more active than the

right (Pulvermuller, 1999b).

Neuropsychological and neurophysiological studies indicate that later-ality of language is present early in life. Infants suffering from brain le-

sions are more likely to develop a temporary language disturbance after

left-hemispheric lesion than after a lesion to the right hemisphere (Woods,

1983). The great plasticity of the neural substrate allows for recovery in mostcases of early neurological language impairment. EEG recordings in young

infants demonstrated a physiological correlate of language laterality early

Page 58: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 58/331

3.2 Laterality of Language 41

in life, even in the first year (Dehaene-Lambertz & Dehaene, 1994; Molfese,

1972).

In which way does the laterality of language functions relate to struc-tural asymmetries? Numerous putative anatomical correlates of languagelaterality have been reported, even in craniofacial asymmetries during early

ontogenetic stages (Previc, 1991). Neuroanatomical Language laterality was

proposed to be reflected neuroanatomically in the size of language-relevant

areas (Geschwind & Levitsky, 1968; Steinmetz et al., 1991), and the size

(Hayes & Lewis, 1993), ordering (Seldon, 1985), local within-area connec-

tions (Galuske et al., 2000), and dendritic arborization patterns (Jacobs et al.,

1993; Scheibel et al., 1985) of cortical pyramidal neurons. These anatomicaldifferences may have a causal role in determining which hemisphere be-

comes more important for processing spoken language, although the causal

chains are, as mentioned, not yet understood. However, one may well argue

that some of the structural asymmetries are a consequence of functional dif-

ferences – for example, of more strongly correlated neuronal activity in therespective areas (see Chapter 2).

Considering the documented anatomical and functional asymmetries, it

becomes important to illuminate the possible causes of left hemispheric lat-erality of language, and to think about the actual causal chain. According

to one view, specific neuroanatomical differences between the hemispheres

cause laterality of neurophysiological processes important for distinguish-ing language sounds, or phonemes. Based on an extensive review of the

neuroanatomical literature, Miller (1996) found that the ratio of white to

gray matter volume yields a smaller value for the left hemisphere in compar-

ison to the right hemisphere. The structures most crucial for language, the

frontal and temporal lobes, exhibit a smaller volume of white matter in theleft hemisphere. Thus, a smaller white matter volume appears to be relatedto language dominance. The white matter is made up primarily of axons

and their glia sheaths, the long-distance cables connecting cortical neurons.

A smaller white matter volume may indicate that average cortical connec-

tions are thinner, and this implies that these connections conduct action

potentials more slowly (Lee et al., 1986). This line of thought leads Miller to

propose that the left hemisphere includes more slowly conducting fibers. In

local cortical circuits, slow fibers may be advantageous for measuring exacttemporal delays, and measuring exact temporal delays in the order of a few

tens of a millisecond is necessary for making phonemic distinctions such as

between a [t] and a [d]. According to this view, language laterality is phono-

logically related and a direct consequence of neuroanatomical properties of 

the human forebrain. However, this theory is, as all other attempts at furtherexplanation of language laterality, in need of more empirical support.

Page 59: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 59/331

42 From Classic Aphasia Research to Modern Neuroimaging

In summary, laterality of spoken language is supported by neuropsycho-

logical lesion studies as well as neuroanatomical and neuroimaging results.

Although it is a well-established fact, its explanation based on neuroscientificprinciples is tentative. Because of the obvious need for further explaining this

fact on the basis of more fundamental neuroscientific knowledge, it may ap-pear safer to postulate an additional principle underlying cortical language

function, which would need to be added to the neuroscientific principles

outlined in Chapter 2.

(IV) Language processes are lateralized to the dominant hemisphere.

Such a principle must be postulated until the causes of the laterality of language are better understood in terms of more elementary neuroscientific

principles, such as the ones expressed by Principles (I)–(III).Aspointedout,lateralityoflanguagedoesnotimplythatlanguagemecha-

nisms are housed exclusively in the dominant hemisphere. The possibility ex-

ists that the nondominant hemisphere also contributes to language processes

and, furthermore, that it contributes specifically to central aspects of linguis-

tic processing. The imaging studies on language laterality, such as those men-

tioned previously, show that both hemispheres become active during lan-guage tasks, but they do not prove that the contribution of the nondominant

hemisphere to language processing is crucial. However, neuropsychological

work made it possible to show that the right hemisphere alone is capable

of word processing. Studies in patients whose cortical hemispheres were

disconnected (  split-brain patients) or in whom the dominant hemispherewas removed (hemispherectomy patients) to cure an otherwise intract-

able epilepsy showed word-processing abilities with their isolated right non-

dominant hemisphere. The nondominant hemisphere could, for example,reliably distinguish words from meaningless material (Zaidel, 1976, 1985).

This has been shown in so-called lexical decision tasks, in which subjects

must decide whether a written letter string is a word by pressing a button. If the visual stimulus is presented to the left of fixation, the visual information

is being transferred only to the right hemisphere (because the fibers in the

visual pathway cross). Presentation to the right of fixation stimulates the left

hemisphere. Split-brain and hemispherectomy patients were able to perform

well on lexical decision tasks when words were presented in their left visualfield. Thus, the right hemisphere (which was nondominant for language)must have processed them. This proves that, apart from the contribution of 

the nondominant hemisphere to pragmatic aspects of language (Zaidel et al.,

2000), its neuronal circuits are capable of word processing as well.

Thus, the imaging and neuropsychological data mentioned show that the

nondominant cortical hemisphere is activated during language tasks and is

Page 60: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 60/331

3.2 Laterality of Language 43

capable of some language processing on its own. However, one may still hold

the view that the language mechanisms in the nondominant hemisphere do

not become relevant unless there is dysfunction in the main language machin-ery of the dominant half of the cortex. To prove entirely that nondominant

right-hemispheric processes are necessary for unimpaired language process-ing, one must show (i) that additional involvement of the right nondominant

hemisphere improves language processing compared to left-dominant hemi-

spheric processing alone, and (ii) that lesions in the right hemisphere can

reduce language processing abilities.

Can the nondominant hemisphere actually assist the dominant hemi-

sphere in word processing? This question was addressed by lexical deci-sion experiments in which stimulus words and pronounceable pseudo-words

(for example, “moon” vs. “noom”) were presented to the left or right of 

fixation of healthy subjects so that information was being transmitted to

the contralateral cortical hemisphere. As an additional condition, identical

words, that is, two copies of the same word, were shown simultaneously

to the left and right of fixation, so that both hemispheres received the in-formation simultaneously. The results showed clearly that bilateral stim-

ulation led to faster and more reliable processing of words compared tostimulation of the dominant left hemisphere alone. This bilateral advantage

could not be observed, or was substantially reduced, for meaningless letter

strings or pseudo-words, and it was also absent in split-brain patients (Mohr,

Pulvermuller, Mittelstadt, & Rayman, 1996; Mohr, Pulvermuller, Rayman,& Zaidel, 1994; Mohr, Pulvermuller, & Zaidel, 1994; Zaidel & Rayman,

1994). These data are consistent with the view that additional information

processing in the nondominant hemisphere can help the dominant hemi-

sphere process words. The right hemisphere seems to play a role in optimiz-ing word processing (Pulvermuller & Mohr, 1996; Hasbrooke & Chiarello,

1998).Can lesions in the right nondominant hemisphere lead to linguistic im-

pairments? This question was addressed by confronting patients with words

from different categories in a lexical decision task. The result was that pa-

tients with lesions in different parts of the right hemisphere showed selective

degradation of words from specific categories, such as words related to ac-

tions or words related to visual perceptions. These category-specific deficitsrevealed by the lexical decision task show that an intact right nondominant

hemisphere is necessary for unimpaired processing of words from particular

categories (Neininger & Pulvermuller, 2001).

In summary, although language laterality is a well-established pheno-

menon, the nondominant hemisphere contributes to, is sufficient for, and is

also necessary for the optimal processing of language.

Page 61: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 61/331

44 From Classic Aphasia Research to Modern Neuroimaging

3.3 Neuroimaging of Language

The knowledge of the cortical basis of language processes has increasedgreatly. One relevant factor was that new imaging techniques became avail-

able that allowed for visualizing brain activity on much more fine-grained

temporal and spatial scales than was previously possible. The temporal dy-

namics of brain activation can be revealed millisecond by millisecond byneurophysiological imaging techniques such as EEG and MEG. The acti-

vation loci can be revealed using metabolic imaging techniques, positron

emission tomogrophy (PET), and functional magnetic resonance imaging

(fMRI). With fMRI, voxels of cortical volume as small as one cubic mil-limeter can be investigated. Thus, the tools became available to scrutinize

the cortical mechanisms of cognition and language in space and time in themillimeter and millisecond ranges.

Some of the new studies confirmed views resulting from the classical

neurologicalstudiesofaphasia.Forexample,thelanguageareasofBrocaand

Wernicke were found to become active in various language tasks. Language

comprehension experiments revealed strongest activation in the superior

temporal lobe, and speech production tasks were found to elicit strongestactivation in the inferior frontal cortex. Thus, basic predictions of the classic

language model of Lichtheim could be confirmed (Price, 2000). It has also

been suggested that the area most relevant for the processing of meanings

and concepts associated with words is in the left inferior temporal lobe.

As we will see later, the question about the processing of word meaningin the brain is a big challenge for modern neuroimaging research; it con-

stitutes one of the most intensely debated topics in the neuroscience of 

language.The advent of modern imaging research has also led to the generation

of data about language-induced cortical activation that cannot be easily ex-

plained by the Lichtheim model and related classic neurological languagetheories. Broca’s area, the “motor language center” of the Lichtheimscheme,

became active in genuine language perception tasks such as phonetic judg-

ment (Zatorre et al., 1992) or silent reading (Fiez & Petersen, 1998). This

indicates that the function of the anterior “motor” language area is not re-

stricted to language production, but includes aspects of language perceptionas well. Correspondingly, Wernicke’s area, the “auditory language center,”became active during genuine production tasks even when self-perception

of the speech signal was made impossible by adding noise (Paus et al., 1996).

Thus, the posterior “auditory” language area seems to have an additional

role in language production. In conclusion, although the core language ar-

eas were active when language tasks had to be solved, their specific function

Page 62: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 62/331

3.3 Neuroimaging of Language 45

auditory words - fixation verb generation - noun reading

Figure 3.2. Presentation of spoken words led to increased metabolism in perisylvian areas

relative to looking at a fixation cross (Left). In contrast, verb generation activated additionalprefrontal areas and the middle and inferior temporal gyrus. Modified from Fiez, J. A.,Raichle, M. E., Balota, D. A., Tallal, P., & Petersen, S. E. (1996). PET activation of posteriortemporal regions during auditory word presentation and verb generation. Cerebral Cortex ,6, 1–10.

as centers contributing to either speech production or comprehension be-

came questionable. These neuroimaging results agree with neuropsycho-logical findings indicating that both classical language areas are necessary

for both language production and comprehension. Both main language areasappear to become active during and to be necessary for language processing,

although neither of them is sufficient for word comprehension or produc-

tion. Most likely, the two classical language centers play a role in both types

of processes.New experimental paradigms were introduced for imaging studies, be-

cause they were found to induce the most pronounced metabolic or neuro-

physiological changes. An example is the verb generation paradigm in which

subjects are given nouns and must find verbs that refer to actions related tothe referent of the noun. Examples would be “fork” and “eat,” or “horse”

and “ride.” When subjects carried out this task, frontal areas including butnot restricted to Broca’s area and temporal areas overlapping with or adja-

cent to Wernicke’s area were found to become active (Petersen et al., 1989;

Wise et al., 1991).

Figure 3.2 shows the activation pattern in the left hemisphere obtained

in one study in which verbs had to be generated in response to nouns (Fiez

et al., 1996). When words had to be listened to, there was activity in superiortemporal and inferior frontal areas that was absent when subjects had to

fixate a cross. However, when new words had to be generated, there was

much more widespread activation in the frontal and temporal lobes. This

additional activity was also present when word generation was compared

with reading of words. This result was used to argue that word processing

Page 63: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 63/331

46 From Classic Aphasia Research to Modern Neuroimaging

can relate to the activation of various areas, not just to the classical language

areas of Broca and Wernicke. Although this conclusion is compromised

by difficulties in interpreting results from word generation experiments, itreceived support from additional neuroimaging work (for discussion, see

Pulvermuller, 1999).Which areas are relevant for processing which aspectof language? Clearly,

if one must generate a word, this makes one think about possible candidates

and alternatives and their respective adequacy, advantages, and disadvan-

tages. If one must perform a more automatized task such as reading words

or even reading very common words, this requires much less effort; con-

sequently, some of the differences in brain activation between tasks canbe accounted for easily in terms of the difficulty of the experiment. Diffi-

culty is an ambiguous word here, and it may be better to speak about the

degree of focused attention required; the amount of attention that must be

shared between different cognitive entities (for example, a presented word

and a to-be-generated word), the amount of effort that must be invested in

a search (for example, for a new item), and numerous other cognitive pro-cesses that can become relevant in language-related tasks, including keep-

ing items in memory, comparing items with each other, and rejecting items.Apart from these aspects, it is not surprising that the modalities involved in

language tasks are related to the activation pattern. For example, a task in

which visual stimuli are to be read and a task in which auditory stimuli must

be processed may more strongly activate the superior temporal or poste-rior occipital lobe, respectively. A task involving language production may,

compared to a mere comprehension task, induce more neuronal activity in

action-related areas in the frontal lobe. Apart from such more or less trivial

differences that can be expected to occur, and actually do occur, betweenlanguage tasks, one may ask whether there are more specific correlates of 

language processing, for example, of the processing of the meaning of a word,its semantics.

The question concerning the cortical locus of the processing of word

meaning has been addressed in many imaging studies, and a number of 

conclusions have been proposed. In a discussion in Behavioral and Brain

Sciences, researchers involved in neuroimaging and neuropsychological re-

search summarized their opinions about the cortical locus of semantic pro-cesses. Figure 3.3 presents the outcome of this discussion. Based on PET

and EEG data, Posner and DiGirolamo (1999) argued for semantic pro-

cesses in left inferior frontal areas. Salmelin, Helenius, and Kuukka (1999)

reported MEG evidence that the left superior temporal lobe is relevant

for word semantics. In Tranel and Damasio’s (1999) framework, the role of the left inferior and middle temporal gyri was emphasized on the basis of 

Page 64: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 64/331

3.3 Neuroimaging of Language 47

Figure 3.3. Cortical areas proposed to be related to the processing of word meaning. Thename of the authors postulating the relevance of the respective areas is printed on top of Brodmann’s area map. Note that different authors attribute the processing of word-relatedsemantic information to different cortical areas. Modified from Pulverm ¨ uller, F. (1999). Wordsin the brain’s language. Behavioral and Brain Sciences, 22 , 253–336.

both lesion evidence and PET data. Skrandies (1999) reported EEG stud-

ies suggesting that the occipital lobes might be involved in distinguishing

word meanings. Epstein (1999) referred to the neuropsychological modelby Geschwind (1965a, 1965b), according to which the angular gyrus at the

boundary of the parietal, temporal, and occipital lobe relates to semantics.

Finally, the opinion was expressed that action-related areas – primary motor,premotor, and prefrontal areas – also may be relevant for aspects of word-

meaning processing and representation (Pulvermuller, 1999a). Regarding

word semantics, Posner and DiGirolamo’s (1999) statement “there is somedispute about the exact areas involved” is clearly correct. As this discussion

shows, modern imaging work has not always led to a satisfactory answer of 

long-standing and important questions about the brain–language relation-

ship. In spite of the undeniable progress made possible by empirical results

obtained with newly introduced techniques, it is likely that theoretical ad-vances are necessary as well. These may help to explain why the processingof meaning – and other language and cognitive processes – should activate

certain cortical areas but not others, and may, in the best case, help us un-

derstand the diversity of results reported in some areas of the cognitive

neuroscience of language. So far, it appears that many imaging studies sug-

gest that numerous areas outside the classical language areas can become

Page 65: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 65/331

48 From Classic Aphasia Research to Modern Neuroimaging

Figure 3.4. Freud’s proposal of how a word is organized in the human brain. Multimodalassociations are thought to exist between neuronal elements housed in different corticalsystems. This model was the seed for modern theories of language processing in the brain(see Chapter 4). Adopted from Freud, S. (1891). ZurAuffassung derAphasien. Leipzig, Wien:

Franz Deuticke.

relevant for the processing of words and their meanings. This important issue

is discussed further in Chapters 4 and 5.Many areas outside the classical language areas of Broca and Wernicke are

relevant for language processing. This idea is suggested not only by modern

imaging research, but actually has a long history. In his monograph on apha-

sia, Freud (1891) proposed a neurological language model that competes

with the one put forward by Lichtheim. Word representations are viewedas multimodal associative networks distributed over various cortical areas(Fig. 3.4). Unfortunately, this proposal was not well received over 100 years

ago, a fact that forced Freud into a career outside the neurological domain.

However, his ideas appear modern and attractive today. As shown in sub-

sequent chapters (e.g., 4–6), models developed on their basis can explain

important results from neuropsychological and neuroimaging research on

language.

3.4 Summary

In the light of neuropsychological and neuroimaging research, Lichtheim’sconnectionist model (Fig. 3.1) was partly confirmed. However, it also became

apparent that this model is too crude to explain many patterns of brain acti-

vation induced by language processing and the more fine-grained aspects of 

language disorders caused by brain lesions. It appears likelythat the language

centers of Broca and Wernicke are mutually functionally dependent. Fur-thermore, although these core language areas are certainly important for lan-guage, they do not appear to represent the only cortical areas contributing to

and necessary for language processing. This is demonstrated by neuroimag-

ing research showing that, in addition to the core language areas, other areas

become active when specific language stimuli are being processed, and by

converging neuropsychological reports about patients with lesions outside

Page 66: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 66/331

3.4 Summary 49

their core language areas who showed category-specific linguistic deficits.

These studies demonstrate that, in addition to the core language areas of 

Broca and Wernicke, there are additional orcomplementary language areas

that are activated during language processing whose lesion causes deteri-

oration of aspects of language processing. These are not restricted to thelanguage-dominant hemisphere (the left hemisphere in most right handers),

but include areas in the nondominant right hemisphere as well. For the sake

of clarity, it may appear advantageous to distinguish the core language ar-

eas of Broca and Wernicke in the language-dominant hemisphere from the

complementary language areas in both hemispheres that can also play a role

in aspects of language processing.Itmaybepossibletoexplainthemutualfunctionaldependenceofthecore

language areas and the category-specific role of complementary language ar-

eas. In Chapter 4, a neurobiological model is summarized that attempts at

such explanation. The model uses concepts introduced in Chapter 2 – namely,

the idea concerning functional cortical webs for words and other language

elements. These are proposed to be cortically organized as strongly con-nected assemblies of neurons whose cortical distributions varies with word

type. For example, concrete words referring to objects and actions are pro-posed to be organized as widely distributed cell assemblies composed of neu-

rons in sensory and motor areas involved in processing the words’ meanings

(see Chapter 4). In contrast, highly abstract grammatical function words and

grammatical affixes are assumed to be more focally represented in the left-hemispheric core language areas of Broca and Wernicke (see Chapter 6).

This proposal builds upon the classic model, because neuron sets in core

language areas are believed to be relevant for all types of language-related

processes. However, it offers a dynamic perspective on the “concept areaB” (Fig. 3.1) by proposing that complementary neurons in areas related to

actions and perceptions regularly involved in language use contribute tocategory-specific language processes.

Page 67: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 67/331

CHAPTER FOUR

 Words in the Brain

This chapter complements Chapter 3 in providing neuroimaging and neu-

ropsychological data about language. Here, the focus is on words. It is asked

which brain areas become active during, and are relevant for, the processing

of words in general, and that of specific word categories in particular.

An aim of this chapter is to show that the neuroscientific principles dis-

cussed in Chapters 2 and 3 give rise to new ideas about the representation andprocessing of words in the brain. The cortex, a neuroanatomically defined

associative memory obeying the correlation learning principle, allows for the

formation of distributed functional webs. During language acquisition, the

neurobiological principles governing the cortex interact to yield the neuron

machinery underlying language. Distributed functionally coupled neuronal

assemblies, functional webs, are proposed to represent meaningful language

units. These distributed but functionally coupled neuronal units are proposed

to exhibit different topographies. Their cortical distribution is proposed torelate to word properties. It is asked how this idea fits into evidence collected

with modern neuroimaging techniques.

4.1 Word-Form Webs

Early babbling and word production are likely caused by neuron activity in

cortical areas in the inferior frontal lobe, including the inferior motor cortex

and adjacent prefrontal areas. The articulations cause sounds, which activateneurons in the auditory system, including areas in the superior temporal lobe.

The fiber bundles between the inferior frontal and superior temporal areas

provide the substrate for associative learning between neurons controlling

specific speech motor programs and neurons in the auditory cortical sys-

tem stimulated by the self-produced language sounds. The correlation learn-

ing principle implies the formation of such specific associations resulting in

50

Page 68: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 68/331

4.1 Word-Form Webs 51

functional webs distributed over the perisylvian cortex, which includes the

inferior frontal and superior temporal core language areas. Figure 4.1a in-

dicates the approximate left-hemispheric distribution of a functional webenvisaged to realize a phonological word form. If neurons in the dominant

left hemisphere are more likely to respond specifically to phonological fea-tures in the acoustic input, the resulting phonological networks must be

lateralized, in the sense of having more neurons in one hemisphere than in

the other. These lateralized perisylvian neuron ensembles would later pro-

vide the machinery necessary for activating a word’s articulatory program

as a consequence of acoustic stimulation with the same word form. This is

necessary for the ability to repeat words spoken by others.Interestingly, babbling, the infant’s earliest language-like articulations,

starts around the sixth month of life (Locke, 1989) and is immediately fol-

lowed by the development of electrophysiological indicators of memory

traces for phonemes (Cheour et al., 1998; Naatanen et al., 1997) and the

infant’s ability to repeat words spoken by others (Locke, 1993). These ob-

servations are consistent with the idea that babbling is essential for buildingup language-specific neuronal representations, in particular sensori-motor

links that may in turn be essential for the ability of spoken word repetition.Word production, in the context of repetition or otherwise, may be essential

for the build-up of specific neuronal representations of individual words.

It might be considered a shortcoming of this proposal that only a minor-

ity of word forms are learned by the infant based on single-word repetition(Pulvermuller, 1999b). How would infants know about which phonemes

actually belong to one word or morpheme if it is spoken in continuous utter-

ances of several words, with many word boundaries unmarked by acoustic

cues? The answer is again implied by the correlation learning principle. Therecurring sound sequences constituting words can be distinguished on statis-

tical grounds from the more accidental sound sequences across word bound-aries (Brent & Cartwright, 1996; Harris, 1955; Redlich, 1993). Behavioral ev-

idence suggests that young infants distinguish the correlated phoneme and

syllable sequences making up words from more accidental sound sequences

in their acoustic input (Saffran, Aslin, & Newport, 1996). Therefore, it ap-

pears likely that single word input is not necessary to build up word repre-

sentations, but that infants can use the correlation statistics, the transitionalprobabilities and/or mutual information of phoneme and syllable sequences

(Shannon & Weaver, 1949), for learning words from continuous speech.

After an auditory word representation has been established by correlation

learning, the repeated articulation of the word made possible by the sen-

sorimotor links set up by babbling would finally establish the word-related

functional web.

Page 69: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 69/331

Figure 4.1. (a) The functional webs realizing phonological word forms may be distributedover the perisylvian areas of the dominant left hemisphere. Circles represent local neuronclusters and lines represent reciprocal connections between them. (b) Word presentationinduced stronger gamma-band responses in the 30 Hz range compared to pseudo-wordpresentation, in particular over the left hemisphere. Reverberatory circuits within wordwebs may underlie the enhancement of high-frequency responses to words compared topseudo-words. (c) The magnetic correlate of the Mismatch Negativity, the MMNm, wasstronger in response to words compared to pseudo-words. Significant differences appearedalready around 150 ms after the word recognition point, suggesting that the activation of word-related functional webs (lexical access) is an early process. (d) The main generator of the word-evoked magnetic mismatch response was localized in the left superior temporallobe. Adopted from Pulverm ¨ uller, F. (2001). “Brain reflections of words and their meaning.”Trends in Cognitive Sciences, 5, 517–24.

Page 70: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 70/331

4.1 Word-Form Webs 53

How would it be possible to prove the existence of functional webs rele-

vant for the processing of words? One view on the nature of functional webs

is that their massive reverberatory circuits produce precisely timed high-frequency rhythms when active (Pulvermuller, Birbaumer, Lutzenberger,

& Mohr, 1997; Milner, 1974; Tallon-Baudry & Bertrand, 1999; von derMalsburg, 1995). Based on this assumption, the prediction is that words

in the input activate the corresponding functional webs, thereby eliciting

strong high-frequency rhythms. In contrast, phonologically and orthograph-

ically regular pseudo-words that are not part of the language would fail to

activate a corresponding functional web, and the high-frequency activity in

the perisylvian areas should therefore be relatively low.This prediction was put to a test using MEG. In fact, a frequency band

around 30 Hz revealed significant differences between acousticallypresented

words and pseudo-words. About one-half second after the onset of spo-

ken one-syllable words, high-frequency brain responses were significantly

stronger compared to the same interval following pseudo-words. Figure 4.1b

shows the results of spectral analyses carried out on data recorded closeto the left anterior perisylvian areas and the homotopic areas in the right

hemisphere of a subject listening to words and pseudo-words. Word-inducedhigh-frequency responses were markedly stronger compared to pseudo-

word-induced activity, both in the single subject whose data are displayed

(12 percent) and in the group average (8.4 percent) (Pulvermuller, Eulitz

et al., 1996). This cannot be due to a global enhancement of the signal be-cause event-related magnetic fields tended to be weaker for words than for

pseudowords in the time window analyzed. EEG- and MEG-studies con-

firmed that words evoke stronger high-frequency brain activity than compa-

rable wordlike material (Eulitz et al., 2000; Krause et al., 1998; Lutzenberger,Pulvermuller, & Birbaumer, 1994; Pulvermuller, Preissl, Lutzenberger,

& Birbaumer, 1996). An explanation of these results can be based on theassumption that word presentation activates functional webs, including mul-

tiple reverberatory circuits, that fail to become fully active if pseudo-words

are perceived.

Physiological differences between words and pseudo-words have been

found in numerous studies using both electrophysiological and metabolic

neuroimaging techniques (Creutzfeldt, Ojemann, & Lettich, 1989; Diesch,Biermann, & Luce, 1998; Hagoort et al., 1999; Price, Wise, & Frackowiak,

1996; Rugg, 1983). Thus, it is uncontroversial that the brain distinguishes

between words and similar but novel and meaningless items. It is, how-

ever, unclear whether such physiological distinction would necessitate that

experiment participants focus their attention on features of the stimuli orengage in language-related tasks. A further question is at which point in time,

Page 71: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 71/331

54 Words in the Brain

after the information about a word is present in the input, the brain makes

the word–pseudoword distinction, so to speak. If distributed functional webs

underlie word processing, an incoming verbal stimulus should automati-cally activate its corresponding representation. Given a sufficient number

of input units specializing, for example, in the detection of acoustic stimulusfeatures are activated, the entire strongly connected web would ignite au-

tomatically as a result of the strong feedforward and feedback connections

holding the network together. This process of ignition (Braitenberg, 1978a)

of the functional web should take place very rapidly, the major factor deter-

mining the latency being axonal conduction delays and temporal summation

of neuronal activity. Axons can bridge large distances in the cortex withina few milliseconds, and the most common corticocortical fibers that havediameters of 0.5–1µm can be estimated to propagate action potentials within

10–20 ms over distances of 10 cm (Aboitiz et al., 1992). A strongly con-

nected distributed neuron web should therefore become active shortly after

its initial stimulation, certainly within 100–200 ms, to use a conservative

estimate.

Neurophysiological recordings, rather than the much slower metabolic

neuroimaging techniques, are necessary to determine the point in time whenthe brain distinguishes words from pseudo-words. Some studies such as

the studies of high-frequency cortical responses discussed previously have

indicated that word-related brain processes can be detected late, that is,

around 400 ms after the presence of the relevant information in the input.However, physiological word–pseudo-word differences in the event-related

potential (ERP) of the brain have been found substantially earlier, in the

so-called N1–P2 complex, 100–200 ms after onset of visually presented

stimuli (Rugg, 1983).In one series of EEG and MEG studies, we could confirm the early pres-

ence of neurophysiological indicators of word and pseudo-word processing.The Mismatch Negativity (MMN) and its magnetic equivalent (MMNm),

which can be elicited by rare changes in the acoustic environment, were

used. The MMN and MMNm were chosen because they have been found

to reflect the existence of memory traces or engrams in the cortex and be-

cause they are largely independent of the subject’s attention (Naatanen,

2001; Naatanen & Winkler, 1999). In earlier research, the MMN had beenfound to reflect the presence of memory traces for phonemes of the subjects’

mother tongue (Naatanen, 2001). In a recent series of EEG and MEG stu-

dies, the neurophysiological correlates of spoken words were compared with

the activity elicited by meaningless pseudo-words (Korpilahti et al., 2001;

Pulvermuller, Kujala et al., 2001; Shtyrov & Pulvermuller, 2002b). To con-

trol for the physical difference, which necessarily distinguishes any word

Page 72: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 72/331

4.1 Word-Form Webs 55

from a pseudo-word, two-syllabic items ending in the same second syllable

were chosen. Between the two syllables was the pause characteristic of some

Finnish consonants, so-called double-stop consonants (for example, “kk”).This pause made it possible to record separate non-overlapping brain re-

sponses to the two individual syllables of a naturally spoken bisyllabic word.To give an example, the Finnish word “pakko” meaning “compulsion” was

contrasted to the meaningless pseudo-word “takko,” and physically identi-

cal “ko” syllables were used after the pause seperating the “pa” or “ta” on

one side and the critical syllable “ko” on the other. Thus, the same critical

syllable was placed in a word or pseudo-word context, respectively.

When the critical second syllable completed a word, its MMN and MMNmwere larger compared to when the syllable was placed in a pseudoword con-

text (Fig. 4.1c). The respective difference was most pronounced 100–200

ms after the word recognition point of the lexical items. The word recogni-

tion point  is the earliest point in time when the information present in the

acoustic input allows the subject to identify the word with some confidence

(Marslen-Wilson & Tyler, 1980). This suggests that the functional web activ-ated by a word in the input becomes active quite early. This finding is also

consistent with recent results from dynamic statistical parametric mappingbased on fMRI and MEG (Dale et al., 2000). These results indicate access

of the cortical representation of words within the first 200 ms after stimulus

information is present in the input. (Pulvermuller, Kujala et al., 2001).

The main source of the cortical generator of the MMNm was localizedin the left superior temporal lobe (Figure 4.1d). Whereas this source was

stronger for words than pseudo-words, its anatomical locus did not change

with lexical status.

It thus appears that the brain can distinguish words from pseudo-words quite early after the relevant information is present in the input.

Still, other physiological word–pseudoword differences – in particular, inhigh-frequency activity – tend to occur with longer latencies. This may, in

part, be related to differences between experiments (regarding stimuli, tasks,

etc.), but may as well indicate that different brain processes are reflected by

these measures. Early ERP differences may reflect the initial full activation,

ignition (Braitenberg, 1978a) of memory traces for words, a putative corre-

late of word recognition, whereas differences in high-frequency responsesmay reflect continuous reverberatory activity of word-related functional net-

works, a putative state of active memory (Fuster, 1997).

It is noteworthy that in the studies of the MMN and MMNm elicited

by words (Korpilahti et al., 2001; Pulvermuller, Kujala et al., 2001; Shtyrov

& Pulvermuller, 2002a, 2002b), the early enhancement of these responses to

words was seen, although experiment participants were instructed to direct

Page 73: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 73/331

56 Words in the Brain

their attention away from the acoustic input and watch a silent movie. To-

gether with results from metabolic imaging studies (Price et al., 1996), the

physiological distinction of words and pseudo-words in these experimentsproves that attention to words is not necessary for activating the words’

cortical memory traces.In summary, physiological studies provide support for the existence of 

wordrepresentationsinthebrain.Theenhancedhigh-frequencyresponsesin

the gamma band to words are consistent with coordinated fast reverberatory

neuronal activity generated by functional webs.

4.2 Category-Specific Word Webs

Word use in the context of objects and actions leads to associations betweenneurons in the cortical core language areas and additional neurons in areas

processing information about the words’ referents. This is implied by the

correlation learning principle and the cortex’s long-range connections be-

tween motor and sensory systems. Functional webs could therefore provide

the basis for the association, in the psychological sense, between an animal

name and the visual image it relates to, or between an action verb and theaction it normally expresses. Strong links within the web can account for

one’s impression that the image is automatically aroused by the word form

presented alone, and that, vice versa, the image almost automatically calls

the name into active memory. The neuron ensembles linking phonological

information and information about the actions and perceptions to which a

word refers are termed word webs here. They would include the phonologi-cal webs in perisylvian areas and, in addition, neurons in more widespread

cortical areas critically involved in processing perceptions and actions, and,additional neurons in various cortical sites where sensory and action-related

information converges and is being merged. The type of entity a word usu-

ally refers to should be reflected in the cortical topography of the functional

web that realizes it.

4.2.1 Visually Related and Action Words

The meaning of an animal name, such as “whale” or “shark,” is usuallyknown from visual experiences, pictures, or films, whereas the meaning of 

a tool name, such as “nail” or “fork,” refers to objects one uses for certain

actions. This is not to say that one could not know a whale from interact-

ing with it or forks from looking at them, but it may appear plausible thatin most individuals more relevant information characterizing whales and

Page 74: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 74/331

4.2 Category-Specific Word Webs 57

forks is visually or action related, respectively. In principle, to draw firm

conclusions on perceptual and functional attributes of word and conceptual

categories, the perceptual and action associations of the stimuli must be eval-uated empirically. The lack of such stimulus evaluation is a caveat of many

studies of category-specific brain processes. Behavioral investigations car-ried out with healthy volunteers revealed that many animal and tool names

show the expected differential elicitation of visual or action associations,

respectively. However, the most striking double dissociation in perceptual

and action attributes was seen between action verbs on the one hand and se-

lected nouns referring to animals or large human-made objects on the other

(Fig. 4.2d) (Pulvermuller, Lutzenberger, & Preissl, 1999). In addition, cat-egories such as “animal names” were not well defined with regard to themodality for which most striking associations were being reported. For ex-

ample, whereas words such as “whale” or “shark” are reported to elicit

primarily visual associations, the results for “cat” are less clearcut, for ob-

vious reasons. Thus, the differential associations cut across the categories

suggested by a philosophical approach (e.g., living vs. nonliving), as was ear-lier found for category-specific neuropsychological deficits (Warrington &

McCarthy, 1987). The sensory/action modalities through which the referentof a word is known appear to be relevant (Fuster, 1999).

It has been argued that it is a possible limitation of this line of thought

that it can be applied only to communication, where words are being learned

in the context of their referent objects or actions. However, word meaningscan also be picked up from contexts in which the actual referents are absent.

Their meaning is frequently revealed by other words used in the same sen-

tence or piece of discourse. It has been proposed that word meaning can be

defined in terms of the set of other words that frequently co-occur with agiven word (Landauer & Dumais, 1997). This would translate into a different

neurobiological scenario for the learning of word meaning. Given that thereis a stock of words whose meaning has been acquired based on word–object

or word–action contingencies, a new word occurring in good correlation with

such known words would only activate its phonological perisylvian repre-

sentation, because no semantic (reference related) links have been set up.

However, neurons in extraperisylvian space related to the known mean-

ing of common context words would frequently be active together with thephonological web of the new word. The correlated activity of semantically

related neurons included in the neuronal representations of known context

words and the phonological web of the new word may allow for “piggy

bay” learning of word meaning. Clearly, this implies that the semantically

related neurons will finally be shared between previously known and new

Page 75: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 75/331

58 Words in the Brain

Figure 4.2. (a) Visual and action associations of words may be mapped by functional websextending over perisylvian language areas and additional visual- and action-related ar-eas in the temporo-occipital and fronto-central areas. The cortical topography of word-related functional webs of words primarily characterized by visual associations may thereforediffer from those of words with strong action associations. (b) Differences in metabolic brainactivation related to the processing of nouns referring to animals and tools in a naming

task. Whereas the tool words more strongly activated a premotor region and an area in themiddle temporal gyrus, animal names most strongly aroused occipital areas. (c) Electrophys-iological differences between nouns and verbs in a lexical decision task recorded at central(close to motor cortex) and posterior (above visual cortex) recording sites. Gamma-bandresponses in the 30 Hz range were stronger close to the motor cortex for action verbs, andstronger above visual areas for nouns with strong visual associations. A similar difference wasrevealed by event-related potentials submitted to Current Source Density Analysis (CSDA).(d) Behavioral experiments showed that the stimulus nouns elicited strong visual associa-tions whereas the verbs were primarily action related. Adopted from Pulverm ¨ uller, F. (2001).“Brain reflections of words and their meaning.” Trends in Cognitive Sciences, 5, 517–24.

Page 76: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 76/331

4.2 Category-Specific Word Webs 59

words so that their neuronal representations would overlap in their semantic

parts. This line of thought shows that the learning of word meaning based on

correlated neuronal activity is, by no means, restricted to the word–objectcontingency scenario. Word meaning can also be learned from context.

Figure 4.2a sketches the postulated neuronal architectures of functionalwebs representing words with strong visual vs. action associations, respec-

tively. The circles in the diagrams symbolize local clusters of neurons strongly

linked by corticocortical fibers. The diagrams illustrate the idea of word webs

that include neurons reflecting semantic aspects. More precisely, the proposal

is that cortical neurons processing aspects of the words’ typical referents, the

entities they refer to, are woven into the networks. If the referent is an objectusually perceived through the visual modality, neurons in temporo-occipitalareas should be included in the web. If a word refers to actions or objects that

are being manipulated frequently, neurons in fronto-central action-related

areas are assumed to be wired into the cortical representations. This can eas-

ily be extended to other sensory modalities as well (Pulvermuller, 1999b). A

shortcoming of the diagrams is that only one type of association is shown for

each word web. Usually, a word that is primarily visually related is reported

to elicit some action associations as well; conversely, an action-related wordyields some visual associations (cf. Fig. 4.2d). The all-or-nothing aspect one

may infer from the diagrams is therefore unrealistic. To make the diagrams

more realistic, each web should include some additional neurons in the re-

spective other field, although the ensembles’ neuron density in these addi-tional areas would be lower than in the areas processing the words’ primary

referential aspect (Kiefer & Spitzer, 2001; Pulvermuller, 2001). Further-

more, action associations imply that there are associations to self-perceived

aspects of the action in the somator-sensory or visual modality. The visualperception of one’s own hand during knocking or writing likely arouses

neurons in movement-related areas of the visual system not activated if, forexample, a stationary visual stimulus is perceived. Therefore, there would

be good reason to add detail to the diagrams, however, at the cost of making

them more complex. Nevertheless, the topographies of functional webs may

differ between semantic word types, and the diagrams may convey the gist

of this idea. (Fig. 4.2a, Fig. 4.3a).

The postulated differential topographies of word webs imply meaning-related processing differences between word categories. A major source of 

evidence for such differences is neuropsychological patient studies in which,

for example, the production or comprehension of nouns and verbs or ani-

mal and tool names was found to be affected differentially by disease of the

brain (Bak et al., 2001; Cappa et al., 1998; Daniele et al., 1994; Miceli, Silveri,

Nocentini, & Caramazza, 1988; Miceli, Silveri, Villa, & Caramazza, 1984;

Page 77: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 77/331

60 Words in the Brain

Patterson & Hodges, 2001; Warrington & McCarthy, 1983; Warrington &

Shallice, 1984; see also Chapter 3). These dissociations between kinds of 

words and conceptual categories can be understood on the basis of the as-sumption of distributed neuron ensembles reflecting perceptual and struc-

tural attributes, including visual features and the degree of overlap betweenexemplars, and the functional attributes, the actions to which the words and

concepts relate (Humphreys & Forde, 2001).

It also may be asked whether the intact human brain demonstrates differ-

ential activation of brain areas when action- or perceptually related words

are being processed. One example of a critical prediction is, if words of one

kind are characterized by stronger action (visual) associations than are wordsof another kind, their processing should be accompanied by stronger brain

activity in the relevant action- (sensory-) related areas. Relevant action-

related areas are in the frontal lobe; the areas necessary for object perception

are in the occipital and inferior temporal lobes.

When pictures of animals and tools were presented in a naming experi-

ment, several areas, including occipital and temporal sites and the classicallanguage areas, were found to increase their activity (Martin et al., 1996).

Category-specific activation was found in the premotor cortex and the mid-dle temporal gyrus when tools had to be named silently, and in the occipital

and inferior temporal lobe when animals had to be named (Fig. 4.2b). These

results meet the previously noted predictions. One may speculate that the

premotor activation is related to the action associations of tool names, asthe activation in inferior–temporal and occipital areas may be related to

visual attributes of animal names. The additional activation in the middle

temporal gyrus in tool naming may be related to movement associations

elicited by the words involved. Differential cortical activation by action-and visually related concepts and words were confirmed, in part, by more re-

cent metabolic imaging studies of category-specific processes using PET andfMRI (Damasio et al., 1996; Grabowski, Damasio, & Damasio, 1998; Martin

& Chao, 2001; Moore & Price, 1999; Mummery et al., 1998; Noppeney

& Price, 2002; Perani et al., 1999; Spitzer et al., 1998; Warburton et al.,

1996),althoughnotallresearchercouldconfirmsuchdifferences(e.g.,Devlin

et al., 2002).

Neurophysiological investigation of noun and verb processing providedfurther evidence for category-specific brain processes relevant for language

(Brown & Lehmann, 1979; Koenig & Lehmann, 1996; Molfese, Burger-

Judisch, Gill, Golinkoff, & Hirsch-Pasek, 1996; Preissl, Pulvermuller,

Lutzenberger, & Birbaumer, 1995; Pulvermuller, Lutzenberger et al., 1999;

Pulvermuller, Preissl et al., 1996; Federmeier et al., 2000). In one study(Pulvermuller, Lutzenberger et al., 1999), differential visual and action

Page 78: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 78/331

4.2 Category-Specific Word Webs 61

associations of the nouns and verbs selected were demonstrated by a rat-

ing study performed by all experiment participants. Event-related potentials

(ERPs) and high-frequency cortical responses revealed a physiological dou-ble dissociation consistent with differential activation of fronto-central and

occipital areas (Fig. 4.2c). The ERP difference was apparent approximately200 ms after the onset of visual word stimuli, consistent with early acti-

vation of the word webs involved. Topographically specific high-frequency

responses, which were stronger over central areas for verbs and over occipi-

tal areas for nouns, started later (400 ms). In an EEG study of auditory word

processing, the physiological distinction between visually related nouns and

action verbs could be replicated, and similar differential activation was foundbetween visually and action-related nouns. In contrast, there was no differ-ence in the topography of brain responses between action verbs and nouns

for whom strong action associations were reported (Pulvermuller, Mohr,

& Schleichert, 1999). These topographical differences in the activation pat-

terns elicited by action-related and visually related words resemble the dif-

ferences observed between written tool and animal names, and between

pictures of animals and tools (Kiefer, 2001). All of these results indicate that

the differential activity pattern evoked by word kinds is not grammaticallyrelated, but rather reflects semantic properties of the stimulus words and

their related concepts. Pulvermuller, Assadollahi, and Elbert (2001) found a

global enhancement of the evoked brain response for a certain subcategory

of nouns, which, according to their behavioral data, had particularly strongsemantic associations to both objects and actions (multimodal semantics).

Control nouns without multimodel semantics failed to elicit the result, again

arguing against an interpretation in terms of grammatical word categories.

Differences in word semantics may also underlie the neurophysiologicaldifferences found between lexically ambiguous and unambiguous words

(Federmeier et al., 2000). This interpretation is suggested because it mayappear plausible that the meaning of ambiguous words is somewhat richer

than that of words with only one meaning. When semantic properties of the

stimulus words were systematically evaluated, we found a linear increase of 

an early component of the event-related magnetic field with a measure of 

the strength of semantic associations of a word (r = 0.8). Therefore, these

data, along with those mentioned earlier, enforce an account in terms of word semantics. It may be that the strong associations, in the psychological

sense, of words with mutimodal semantics are realized as strong connections

within particularly widespread and large cortical neuronal assemblies. Ac-

tivation of these particularly widespread and strongly connected networks

may be reflected as an enhancement of the neurophysiological response

(Pulvermuller, 2001).

Page 79: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 79/331

62 Words in the Brain

4.2.2 Sub-types of Action Words

More fine-grained predictions are possible on the basis of the postulate that

topographies of word webs reflect the words’ referents. Action verbs canrefer to actions performed with the legs (walking), arms (waving), or mouth

(talking). It is well known that the motor cortex is organized somatotopically,

that is, adjacent body muscles are represented in neighboring areas within themotor cortex (He et al., 1993; Penfield & Rassmussen, 1950). Neurons con-

trolling face movements are located in the inferior precentral gyrus, those

involved in hand and arm movements accumulate in its middle part, and

leg movements are controlled by neurons in its dorsomedial portion (see

Chapter 2). The correlation learning principle therefore suggests differen-tial topographies for cell assemblies organizing leg-, arm-, and face-relatedwords (Fig. 4.3a). Differential action-related associations of subcategories of 

verbs could be demonstrated by behavioral studies (Fig. 4.3b; Pulvermuller,

Hummel, & Harle, 2001).

In an EEG study, we compared face- and leg-related action verbs (“walk-

ing” vs. “talking”). Current source density maps revealed differential acti-

vation along the motor strip. Words of the “walking” type evoked stronger

ingoing currents at dorsal sites, over the cortical leg area, whereas thoseof the “talking” type elicited the stronger currents at inferior sites, next to

the motor representation of the face and articulators (Fig. 4.3c; Hauk &

Pulvermuller, 2002; Pulvermuller, Harle, & Hummel, 2000; Pulvermuller,

Hummel, & Harle, 2001). A similar study comparing arm- and leg-related

words was performed using fMRI. The preliminary data shown in Figure 4.3dare consistent with the view that the body parts involved in the actions action

verbs refer to are reflected in the cortical neuron webs these words activate.

Furthermore, the early point in time at which the word category differenceswere present in neurophysiological responses indicates that there was no

substantial delay between word form access (perisylvian activation) and theprocessing of action attributes (more superior activation, for example in

the case of leg words). This supports the view that information about the

word form and the body parts with which the word-related actions are being

carried out, are woven into the same word-related cortical networks and are

activated near-simultaneously (Pulvermuller, 2001).

4.3 The Time Course of Lexical and Semantic Activation

The lexical status of a written or spoken word, whether it is a word or not, and

aspects of word semantics appear to crucially determine the brain response.

The differences between semantic word categories can appear early in the

neurophysiological brain response, that is,∼100–200 ms after stimulus onset

Page 80: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 80/331

4.3 The Time Course of Lexical and Semantic Activation 63

Figure 4.3. (a) Cortical topographies of functional webs representing different types of action words may differ. Action words can refer to actions executed by contracting face,arm, or leg muscles (e.g., to lick, to pick, to kick ). Different neuron ensembles in the primarymotor cortex may therefore be woven into the word-related neuron ensembles (cf. Fig. 1b).(b) Ratings of face, arm, and leg associations confirming differential referential semantics of three action verb groups. (c) Results from an EEG study. Topographical differences betweenbrain responses to face- and leg-related verbs. Stronger ingoing currents were seen closeto the vertex for leg-related items (gray spot at the top) and at left-lateral sites, close to theface representation, for face-related words (dark spot in the middle). (d) Result from a fMRI

study comparing arm- and leg-related verbs (single subject data). Differences were see inthe precentral gyrus of the left hemisphere. Adopted from Pulverm ¨ uller, F. (2001). “Brainreflections of words and their meaning.” Trends in Cognitive Sciences, 5, 517–24.

(e.g., Pulvermuller, Assadollahi et al., 2001; Skrandies, 1998). This latency

range corresponds to the time range where the earliest neurophysiological

differences between words and pseudo-words were found (e.g.,

Pulvermuller, Kujala et al., 2001; Rugg, 1983). Thus, the earliest latencies

at which the lexical status and the semantic category of word stimuli were

Page 81: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 81/331

64 Words in the Brain

reflectedintheneurophysiologicalresponsecoincidedwitheachother.These

neurophysiological data support psycholinguistic models postulating that

information about a word’s meaning can be accessed near-simultaneouslywith information about its form, a proposal motivated by behavioral studies

(Marslen-Wilson & Tyler, 1975; Marslen-Wilson & Tyler, 1980). Likewise,they are consistent with the view that a word is cortically processed by a

discrete functional unit storing information about the word’s form together

with that about its semantics.

While the semantically and form-related parts of distributed word webs

may be activated early and near-simultaneously, there is evidence that dif-

ferent physiological processes occur in sequence in the same cognitive brainrepresentations. A stage of access to the representation (ignition of the cell

assembly, see Braitenberg, 1978a) may be followed by sustained reverber-

atory activity (active memory, see Fuster, 1995) of the word web. Whereas

the early access stage may occur within one quarter of a second after the

information in the input allows for recognizing a stimulus word, the rever-

beratory activity related to active memory would follow after more than250 ms. The early access process may be reflected in early event-related po-

tentials, and the late reverberations may lead to high-frequency responses inthe gamma band. These hypotheses can tentatively explain recent findings

about the time course of neurophysiological responses to words (for further

discussion, see Kiefer, 2001; Pulvermuller, 1999).

4.4 Summary and Conclusions

The brain response to words and word-like material appears to reflect lexical

status and word semantics. Word/pseudo-word and word category differenceswere reported in metabolic and neurophysiological imaging studies. Both

types of differences were found already at 100–200 ms after the information

in the input allowed for recognizing the words, whereas some differences,for example in high-frequency responses, appeared only with longer delays.

These results can be explained on the basis of the idea that words are repre-

sented and processed by distributed but discrete neuron webs with distinct

cortical topographies. They are somewhat less easily explained by alterna-

tive approaches. If words were represented by single neurons, for example,the corresponding specific brain activity states could probably not be distin-guished with large-scale neuroimaging techniques, such as MEG or fMRI.

Also, it is in question how the specific changes observed between words and

pseudowords could be explained if both stimulus types were processed alike

by a distributed network of neurons in which no discrete representations

exist, or by interference patterns over the entire cortex. Furthermore, an

Page 82: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 82/331

4.4 Summary and Conclusions 65

explanation of word-category differences may turn out to be even more dif-

ficult on the basis of such approaches. Thus, while competing approaches are

challenged by the data discussed, the postulate of discrete functional websrepresenting words explains them well.

The results on category differences described here indicate that aspectsof the meaning of words are reflected in the topography of brain activation.

They are also consistent with the view that the referents of particular word

kinds are relevant for determining the brain areas involved in their process-

ing. The data do not explain the entire spectrum of areas found to be active

during category-specific word processing. There are findings about different

semantically related activity patterns that are not readily explained by ele-mentary neuroscientific principles, such as the principles (I) – (IV) discussedearlier. For example, the differential activation of the right- vs. the left pari-

etal areas by names of body parts and numerals (Le Clec’H et al., 2000)

cannot be explained by the four principles alone. It is likely that additional

principles of cortical functioning as yet not fully understood are necessary

to account for these data. Furthermore, it must be added that the semantic

category of the stimulus words is by far not the only variable determin-

ing the topography of brain activation. Clearly, the modality of stimulation(visual or auditory) and the task context in which words must be processed

(e.g., lexical decision, naming, memory) play an additional important role

in determining the set of active brain areas (Angrilli et al., 2000; Mummery

et al., 1998). Moreover, other features of the stimulus material, for example

the length and frequency of words, play an important role (Assadollahi &Pulvermuller, 2001; Rugg, 1990). The present approach suggests, and the

summarized data indicate, that, if possibly confounding variables are appro-

priately controlled for, category-specific differences are present betweenword categories and conceptual kinds across different tasks and stimulus

modalities.

A limitation of the considerations so far is that they are restricted to thelevel of single words. In fact, they apply to word stems that include only one

meaningful unit (morpheme). The neurobiological nature of affixes with

grammatical function, for example, has not been addressed. Also, the neu-

robiological basis of grammatical or function words, for example “if”, “is”,

and “it”, has not been specified, and, of course, the complex interaction be-tween word-related processes during analysis and synthesis of a word chainis a further open issue. Some of these issues will be covered in later chap-

ters. Chapter 5 will address meaning and form relationships between words,

Chapter 6 function words and inflectional affixes, and Chapters 10 to 13

various grammatical phenomena.

Page 83: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 83/331

EXCURSUS ONE

Explaining NeuropsychologicalDouble Dissociations

The cortex may be an associative memory in which correlation learning es-

tablishes discrete distributed functional webs. It is important to ask how this

view relates to clinical observations, in particular to the neuropsychological 

double dissociations seen in aphasic patients. Here is an example of sucha double dissociation. Patient A exhibits severe deficits in producing oral

language (Task 1), but much less difficulty in understanding oral language(Task 2). Patient B, however, presents with the opposite pattern of deficits –

that is, only relatively mild language production deficits, but substantial diffi-

culty in comprehending spoken language. Briefly, Patient A is more impairedon Task 1 than Task 2, whereas Patient B is more impaired on Task 2 than

Task 1.

Clearly, the explanation of neuropsychological syndromes and, in partic-

ular, of double dissociations is an important issue for any model of brain

function and, therefore, a brief excursus may be appropriate here. It wasonce argued that the existence of a double dissociation demonstrates, orstrongly suggests, the presence of modules differentially contributing to

specific aspects of the tasks involved (Shallice, 1988). A module would be

conceptualized as a largely autonomous information processor (see also

Section 6.2). A standard explanation of a double dissociation, therefore,

is that two modules are differentially involved in the two tasks (1 and 2), and

that one of them is selectively damaged in each of the patients (A and B).

The idea of functional cortical webs may lead the reader to conclude thatthe distributed neuronal systems would always deteriorate as a whole when a

lesion takes place, even if only a focal cortical area is affected. The suspicion

may therefore be that double dissociations cannot be explained. This con-

clusion is, however, not justified. Neuropsychological double dissociations

can, in fact, be accounted for by a model composed of distributed functionalwebs.

66

Page 84: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 84/331

E1.1 Functional Changes in a Lesioned Network 67

This excursus explains how a functional web architecture can model a

double dissociation, and why double dissociations do not prove the existence

of modules differentially involved in the tasks on which a double dissociationis observed. The initial paragraphs present considerations and simulations

regarding the influence of small lesions on the function of an ensemble of strongly connected neurons.

E1.1 Functional Changes in a Lesioned Network: The NonlinearDeterioration of Performance with Growing Lesion Size

How does a brain lesion change the function of a strongly connected neu-ronal ensemble? If one neuron in an assembly is lesioned, this may have no

effect at all. However, if a certain percentage of its neurons have been le-sioned or removed, the ensemble becomes unable to serve its functional role

in cerebral life, so that it becomes inappropriate to speak about “full” acti-

vation or ignition when the remaining neurons become active. The smallest

percentage of active neurons (of the intact assembly) necessary for speaking

about an ignition is called the ignition threshold here. In the intact network,

the ignition threshold can be considered the point of no return, the numberof neurons active after which the ignition can no longer be stopped by in-

hibitoryinput.Ifthenumberofassemblyneuronssurvivingalesionissmaller

than the ignition threshold, the ensemble, by definition, cannot ignite after

stimulation. Therefore, it can be considered to be destroyed. Clearly, not

every lesion leads to destruction of all affected ensembles. If the damage ismoderate, the assemblies still ignite after appropriate stimulation, but the

time needed for the ignition to take place is longer. This can be illustrated

by neural network simulations (Pulvermuller & Preissl, 1991).Figure E1.1 gives average ignition times t θ and percentage of destroyed

neuronal ensembles d as a function of the lesion size l. In this simulation, the

ignition threshold has been set to 70 percent of the total number of neuronsof the intact ensembles. Importantly, the assemblies tolerate lesions of a

substantial percentage of their neurons with only minor ignition delays. Even

pronounced lesions (e.g., of 15 percent of the neurons) affect performance

only minimally. However, after a certain critical amount of damaged tissue

has been reached, 20 percent in the presented simulations, performancedeteriorates very rapidly if the lesion size increases slightly.

The nonlinear decline of performance of the networks with increasing

lesion size may have implications for the explanation of progressive neu-

rocognitive impairments. Many progressive brain diseases lead to minor

behavioral deficits, although pronounced structural deficits can already be

observed. Minor additional structural deterioration sometimes leads to rapid

Page 85: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 85/331

68 Explaining Neuropsychological Double Dissociations

Figure E1.1. The effects of lesions (l) of different sizes on neuron ensembles each including100 neurons. The ignition threshold was set to 70% – that is, the assembly was calledignited if 70 of the 100 neurons were active. As a function of lesion size, the average time t θneeded for an ignition, and the percentage d of destroyed assemblies (which could no longerignite) increased. Small lesions did not have a strong effect. In contrast, after removal of approximately 20% of the neurons, further slight increase of the lesion size caused dramaticdysfunction, as reflected by the steep raise of both ignition times t ∅ and the percentage of destroyed ensembles d . Reprinted with permission from Pulverm ¨ uller, F., & Preissl, H. (1991).A cell assembly model of language. Network: Computation in Neural Systems, 2 , 455–68.

Page 86: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 86/331

E1.2 Broca’s vs. Wernicke’s Aphasias 69

and pronounced deterioration of behavior. The simulations show that this is

exactly the way a functional web acts when progressively more of its neurons

are being damaged. Small increases of the lesion size cause pronouncedfunctional impairments only if ignition thresholds are being approached.

E1.2 Broca’s vs. Wernicke’s Aphasias

What is the effect of focal lesions in different parts of the perisylvian lan-

guage cortex in a model of overlapping functional webs distributed over

these cortical areas? Again, the idea is that acoustic, articulatory, and seman-

tic information about words are bound together in functional units exhibitingspecific cortical topographies. This means that these aspects of information

processing are not separate functionally, although, originally, they may pri-marily have been housed in separate brain areas. How would a network of 

several distributed and partly overlapping information processors respond

to focal lesions?

Before associative learning, articulatory programs are controlled by neu-

rons in the prefrontal, premotor, and primary motor cortex, whereas acoustic

properties may activate neurons only in the superior temporal lobe stimu-lated by features of speech sounds, and input related to word semantics

(reference) are possibly exclusively processed in additional brain areas.

After word learning, however, all (or most) of these neurons would be ac-

tivated during comprehension, articulation, and semantic processing of a

word because the functional web has formed. In the same way the neuronsin different areas were functionally separate before learning had taken place,

these distant neuron groups may become functionally separate again after

their strong links have been cut by a lesion, or after one part of the assem-bly has been destroyed. If processes dissociate after the lesion, they may

nevertheless have been linked in the fully functionl brain.

An intact neuronal assembly would include efferent neurons that controlarticulatory movements and afferent neurons stimulated by acoustic phono-

logical input. These neuron groups can be considered to lie in the periphery

of the assembly. In the center are neurons in various association cortices

that do not receive direct input from outside the cortex and do not con-

trol its output directly. Their primary purpose is to bind information. Thesebinding sites can themselves be connected to other areas that do not re-ceive direct input from outside the cortex, or directly control motor output.

Figure E1.2 schematizes a network of several partly overlapping neuronal

ensembles and the way their elements (local neuron clusters) may be dis-

tributed over perisylvian cortices. The network equivalents of areas in the

temporal lobe are represented at the top and those of the inferior frontal

Page 87: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 87/331

70 Explaining Neuropsychological Double Dissociations

Figure E1.2. Structure of a network used for simulating the effect of focal lesions in the peri-sylvian cortex. Assemblies included neurons in primary and higher-order perisylvian areas.Lesions in one of the “peripheral” parts of the assemblies (uppermost input or lowermostoutput layers) led to unimodal processing deficits in the simulation (in either “word produc-tion” or “word perception”). Lesions in the middle (the network equivalent of Broca’s andWernicke’s areas) caused multimodal but asymmetric deficits (e.g., a strong “production”deficit with a moderate “comprehension” deficit, or vice versa). The model therefore ac-counts for a double dissociation of the performance on two language tasks (speech pro-duction vs. comprehension), and, in addition, for the frequent cooccurrence of deficits re-garding these tasks. Reprinted with permission from Pulverm ¨ uller, F., & Preissl, H. (1991).

A cell assembly model of language. Network: Computation in Neural Systems, 2 , 455–68.

cortex are at the bottom. Only the uppermost and lowermost neurons in the

periphery (uppermost or lowermost layers) have efferent or afferent corti-cal connections. Obviously, this is an elementary model based on several

simplifying assumptions.

Word comprehension is modeled in the circuit as stimulation of the upper

auditory input layer and subsequent ignition of one assembly. Word produc-

tion is modeled by stimulation in the center of one assembly followed by itsignition plus activation of its motor output neurons in the lowermost layer.Specific deficits in comprehension or production can now be explained intu-

itively as follows: A lesion in one of the layers or areas of the model destroys

assembly neurons. However, lesions in the periphery, in a primary area, lead

to additional disconnection of the web from its input or output. A moderate

lesion in the uppermost layer may therefore only cause a mild increase of 

Page 88: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 88/331

E1.2 Broca’s vs. Wernicke’s Aphasias 71

ignition times in the simulation of word production, thus leaving word pro-

duction largely ‘intact.’ However, this same lesion may make it impossible

to ignite the assembly through its afferent fibers if word comprehension issimulated. This impossibility does not necessarily imply that all the assem-

bly’s afferent connections have been cut or that all neurons in the auditoryinput layer have been destroyed. Removal of only a few peripheral input

neurons of an assembly can slightly delay its ignition so that its neighbors

may take over and ignite instead of the directly stimulated assembly. This

process corresponds to a failure or error in activating a word-specific neuron

ensemble. One may consider this an implementation of a failure or error in

lexical access.As illustrated by this example, unimodal processing deficits apparentonly in word comprehension can be explained quite naturally in a model

relying on functional webs. A selective deficit in understanding spoken

words whereas other sounds can still be interpreted, has been named word-

  form deafness. What has been said about comprehension can be extended

to speech apraxia on the motor side. Apraxia of speech includes a deficitin producing word forms correctly. Phonological errors, omissions, hesita-

tion phenomena, and other coordination problems in speech productionpredominate, whereas there is no pronounced deficit in word comprehen-

sion. Selective lesion of the motor output layer of the model network in

Figure E1.2 primarily cuts neuron ensembles from their motor output fibers.

This deteriorates the network’s performance on the word production task,

but only minimally shows in simulations of word comprehension. Similar topatients with apraxia of speech, the networks produce omissions and incor-

rect output in the absence of deficits in word comprehension. The simple

model does not allow for modeling fine-grained aspects of apraxia of speechand word-form deafness, but it can explain the important double dissoci-

ation seen between the two types of unimodal syndromes, one specifically

affecting word production and the other word comprehension.The explanation of cases of unimodal deficits, such as apraxia of speech

or word-form deafness, in which two deficits are fully dissociated (each

without the other) is important. It is probably equally important to explain

the fact that symptoms frequently co-occur in certain combinations. Most

aphasias are multimodal deficits involving all language modalities, but eachto a different degree. A simulation may therefore aim at mimicking bothdissociation and co-occurrence of symptoms.

Lesion simulations using cell assembly networks suggest that the closer a

lesion is to the periphery of an assembly, to the primary areas, the more asym-

metrical is the disturbance. As discussed previously, a lesion in the periphery

(i.e., in the input or output layer) of a network, such as the one depicted in

Page 89: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 89/331

72 Explaining Neuropsychological Double Dissociations

Figure E1.2, can lead to a unimodal deficit, either in word comphrehension

or word production. However, lesion in the next layers (second from top

or bottom) was found to cause a bimodal but asymmetric disturbance, forexample, strong comprehension deficit but minor production impairment.

Finally, a lesion in the layers in the middle of the network caused an almostsymmetrical pattern of errors. The model allows for predictions on both the

dissociation and co-occurrence of symptoms caused by focal cortical lesions.

When lesions were allowed in only the upper half of the network, the ana-

log of Wernicke’s area in the superior temporal lobe, the network showed

strongly impaired performance on the simulated word comprehension task

and relatively mild impairments when the network was used to simulateword production. In contrast, a lesion in the lower half of the network – cor-

responding to the inferior frontal region which includes Broca’s area – led to

stronger deficits in the word production task and only milder impairment in

the network’s word comprehension. Thus, the network simulation provides

a model of a double dissociation frequently seen in patients with Broca’s

and Wernicke’s aphasias.Further specific features of Broca’s and Wernicke’s aphasias were also

present in the model. For example, after lesion in the network’s Broca area,activity frequently extinguished in the production task, thus simulating in-

ability to produce a word, which is indeed seen in many patients with Broca’s

aphasia. In contrast, lesions on the Wernicke side of the network rarely pro-

duced such lexical omissions but rather yielded activation of not stimulated(incorrect) ensembles. This nicely corresponds to the frequent use of in-

correct words by patients with Wernicke’s aphasia. (For more details about

these simulations, see Pulvermuller, 1992; Pulvermuller & Preissl, 1991).

These simulations explain one important double dissociation – the differ-ential degradation of word production and comprehension in Broca’s and

Wernicke’s aphasias on the basis of a simple model of functional webs dis-tributed over perisylvian areas. Because individual layers of the network can

be likened to areas in the cortex, the network simulation has implications

regarding the location of cortical lesions that cause particular deficits.

Because a distributed model without functional segregation between

modules was able to imitate a double dissociation pattern, it follows that

no modular structure is necessary for explaining this neuropsychologicalphenomenon. This point has been made earlier (Plaut & Shallice, 1993;

Pulvermuller & Preissl, 1991) and is now widely agreed on. One should,

however, emphasize that these considerations by no means rule out modu-

lar models of cognition; rather, they show that different views are possible

on the cortical cause of double dissociations.

Page 90: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 90/331

Page 91: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 91/331

CHAPTER FIVE

Regulation, Overlap, and Web Tails

Chapter 4 offers a neurobiological perspective on word processing in the

brain. The time course and topography of cortical activation during word

processing, in particular during the processing of words of different cate-

gories, is discussed in some detail. The proposal is that there is a cell ensem-ble or functional web for each and every word, and that words with different

referential meaning may have functional webs characterized by differenttopographies.

Looking at words more closely, more questions arise, for example, regard-

ing complex form–meaning relationship. Two words may share their form

(e.g., “plane,” meaning “aircraft” or “flat surface”), or may sound differentlybut have largely the same meaning (e.g., “car” and “automobile”). There are

word forms that include other word forms (e.g., the letter sequence “nor,”

including “no” and “or”), and there are words whose meaning includes,

so to speak, the meaning of others (e.g., “animal” and “dog”). These rela-tionships of homophony (or polysemy), homonymy (or synonymy), inclu-

sion of word forms in other word forms, and hyperonymy (vs. hyponymy)may pose problems to an approach relying on cortical neuron webs. How

might a neurobiologically realistic model of complex form-meaning relation-

ships look? Tentative answers are discussed in the section on overlap of rep-

resentations.

Other aspects largely ignored in earlier chapters relate to the information

about written language and to aspects of meaning that have been charac-terized as emotional or affective. This information may be related to the

phonological and semantic side of words through connections of the rele-

vant word webs either to subcortical or to additional cortical neurons. A few

remarks are made about the possible circuits in the section on web tails.

A concern comes from considerations on activity dynamics in the cortex.Associative learning networks run into the problemof associating everything

74

Page 92: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 92/331

5.1 Regulation of Cortical Activity 75

with everything if too much learning has taken place, or if too much activity

is allowed to spread. It is argued here that a regulation mechanism is needed

in autoassociative models of network memory and that such a mechanismcan be important for solving the mentioned putative problems regarding

words that are related phonologically or semantically.

5.1 Regulation of Cortical Activity

A brain model in the tradition of Hebb’s cell assembly theory runs into anumber of problems. Milner (1957, 1996) discusses some of these problems

in great detail. An important problem was the result of the original formula-tion of Hebb’s learning rule, according to which two connected neurons that

frequently fire together increase the strength of their wiring. One can call

this a coincidence rule of associative learning because only coincident firingof two neurons is considered to have an effect on connection strength. In a

network with many links between neurons, a coincidence rule can lead to

ever-increasing connection strengths so that, finally, catastrophic overacti-

vation may take place whenever the network is being stimulated.

There are, however, tricks available that allow for solving, or at leastminimizing, the problem of catastrophic overactivation. These tricks seemto be applied by the real brain to allow for effective storage of memories, orengrams, in the cortex. Three strategies for minimizing the overactivation

problem are introduced after this problem itself has been explained in more

detail.

5.1.1 The Overactivation Problem in Auto-associative Memories

The overactivation problem can be illustrated using artificial associative

memory networks. A fully connected autoassociative memory consists of 

n neurons and reciprocal connections among all of them. This network typeis not a particularly realistic model of the cortex, but it can be made more

realistic by omitting connections between neurons, for example by reduc-

ing the number of connections of each neuron to√ 

n (Palm, 1982). In this

case, the network is still called autoassociative, as long as it includes loops.

(A network in which different neuron populations are connected in onlyone direction is called a hetero-associative memory.) The idea underlyingthe Hebbian approach is that the neurons that show frequent coincident

activity, for example, the neurons activated by an object stimulus, link into

a neuronal assembly so that finally there are many object representations in

the network. Each object would be realized as a neuron ensemble with par-

ticularly strong connections between its members. These ensemble–internal

Page 93: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 93/331

76 Regulation, Overlap, and Web Tails

connections are much stronger than the average connection strength in the

entire autoassociative network. Now, the problem is that coincidence learn-

ing leads to an increase in average connection strength. After much learningand strengthening of connections, so many neuronal links may have become

so effective that any stimulation of some of its neurons activates the entirenetwork, a process that may be analogous to an epileptic seizure in the real

nervous system (Braitenberg, 1978a). This is undesirable because the aim

is to keep representations separate and allow for retrieving, or specifically

activating, individual object representations. This property is lost if too many

too strong links developed.

5.1.2 Coincidence vs. Correlation

Obviously, a first strategy to avoid the overactivation problem would be to

avoid the learning rule proposed by Hebb, or at least to modify it. There

is good motivation for this. It has become clear that not only coincidence

of neuronal firing has an effect on synaptic efficacy (Brown et al., 1996;

Rauschecker & Singer, 1979; Tsumoto, 1992). Although coincident activity

of two neurons can lead to synaptic strengthening, antiphasic activation of the neurons can reduce their connection strength. If one neuron fires while

the other is silent, their synapse may become weaker, a process sometimes

called Anti-Hebb learning. The correlation rule mentioned in Chapter 2 cap-

tures important features of the known principles underlying synaptic plas-

ticity, in particular, Hebbian and Anti-Hebb learning. The correlation ruleimplies that each learning event includes strengthening of certain synapses

(the synapses between coactivated neurons) as well as a weakening of other

synapses (the synapses between activated and inactive neurons). Thus, if parameters are chosen appropriately, the average connection strength does

not increase when engrams are being stored in an autoassociative network.

This would make it less likely that, after substantial learning has taken place,the network shows catastrophic overactivation.

5.1.3 Sparse Coding

A second strategy for minimizing the overactivation problem is to use whathas been termed sparse coding. This means that, among the many neuronsin a network, only a small subset of neurons contributes significantly to

the representation and processing of each individual engram stored in the

network. The activity state of the entire network can be coded using a vec-

tor of numbers, with each number giving the activity value of one neuron

of the network. For sparse coding, each engram can be characterized by

Page 94: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 94/331

Page 95: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 95/331

78 Regulation, Overlap, and Web Tails

In sum, sparse coding not only has computational advantages but is also

likely to be used for information storage in the cortex. One of its important

consequences is the generally low activity level in the associative network,which further reduces the likelihood of overexcitation.

5.1.4 A Cybernetic Model of Feedback Regulationof Cortical Activity

Apart from correlation learning and sparse coding, a third strategy is likely

to be used by the brain to prevent overexcitation. Even in an autoassocia-

tive network with moderate average connection strength, strong stimulationfrom outside may overactive the neurons. However, lack of stimulation may

cause activity to die out. To prevent these undesired outcomes, a regulationmechanism designed to keep the activity level within certain bounds must be

postulated. In the real brain, this mechanism would control the activity status

of the cortex. The regulation mechanism would respond to strong activity

increase by a process of inhibition or removal of background activity, that is,

disfacilitation. If activity is about to die out, the control device may provide

more background activity to elevate the excitation level again.An important mechanism for controlling excitation is provided by the

small inhibitory cells in the cortex itself. As mentioned in Chapter 2, local

inhibitory circuits can reduce activity in small pieces of gray cortical matter if 

much excitation is provided by local pyramidal neurons. Thus, overexcitation

in the cortex appears to be counteracted by a local feedback regulationmechanism.

A different more global control mechanism has been formulated in terms

of a cybernetic feedback–regulation circuit (Braitenberg, 1978a). The feed-back regulation device would take the global activity level, A, in the cortex –

or any other part of the brain, B – as its input and compute an output depen-

ding on how far this measured activity value A deviates from a preset targetvalue. The output of the regulation is fed back to the cortex. This feedback

value is called θ , which is calculated, at each point in time, as a function of the

activity state A. It can be conceptualized as global background activity that

is increased or decreased to move A toward the target value. Braitenberg

(1978a) called this mechanism threshold control, a term that may suggestthat the parameter actually changed by the regulation device is the thresh-old of individual cortical neurons. However, as pointed out by Braitenberg,

it is likely that membrane potentials are the variable influenced by feed-

back control in the brain. If the threshold of a neuron is defined in terms of 

the change in membrane potential needed to activate the cell, an increase

in background activity lowers the threshold and a reduction in background

Page 96: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 96/331

5.1 Regulation of Cortical Activity 79

Figure 5.1. The idea of threshold regulation as formulated by Braitenberg. A brain (B) re-ceiving input (I) and producing a functional output (F) has a second output that includesinformation about its activity state (A). Based on this second output, threshold values (θ )of the neurons in the brain are recalculated and fed back to the brain. Feedback control issuch that higher activity in the brain raises the threshold and lower activity attenuates it.A feedback control mechanism is necessary for keeping brain activity within certain limits.Adopted from Braitenberg, V. (1978). Cell assemblies in the cerebral cortex. In R. Heim& G. Palm (Eds.), Theoretical approaches to complex systems. (Lecture notes in biomathe-matics, vol. 21) (pp. 171–88). Berlin: Springer.

activity elevates it again. The threshold control mechanism postulated by

Braitenberg is illustrated in Figure 5.1.

Although there is agreement that regulation of cortical activity is neces-

sary, the exact characteristics of the mechanism and the brain systems that

realize it are still under discussion. For example, the threshold control mech-anism regulating the activity state of an autoassociative network may take

several alternative variables as its input. It could look at the actual number

of neurons active in a given time window or at the average frequency of theiraction potentials. When it comes to specifying the working of the mechanism

in more detail – as shown in Chapters 10–12 on grammar mechanisms – it is

assumed that the threshold control mechanism detects fast activity increasesand provides global inhibition if activity raises substantially. In particular,

the control mechanism is assumed to become active if a previously inactive

neuronal ensemble ignites. Furthermore, the assumption is that excitation

is provided if activity levels are generally low. This is one of several realistic

possibilities to model a threshold regulation mechanism.There are alternative or complementary, proposals highlighting the pos-

sible role of cortico-cortical (Milner, 1996), cortico-striatal (Wickens, 1993),

and hippocampal (Fuster, 1995; Miller, 1991) connections in controlling the

activity level in the neocortex. It is not clear, however, which of these brain

structures is most crucial for regulating the activity level in the neocor-

tex. It may well be that more than one of the well-known loops formed

Page 97: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 97/331

80 Regulation, Overlap, and Web Tails

by the cortex and other brain structures play a crucial role as regulation

devices.

5.1.5 Striatal Regulation of Cortical Activity

One putative circuit is now described in more detail to illustrate how aregulation mechanism might operate. The neocortex is intimately linked to

subcortical brain structures. An important loop is formed by projections

form the cortex to the neostriatum (Putamen and Nucleus candatus), from

there to the paleostriatum (or Pallidum), and finally to the thalamus, fromwhere projects run back to the cortex, in particular the frontal lobes. Two of 

these links are inhibitory, form neostriatum to paleostriatum and from thereto thalamus, so that the two inhibitory connections in series produce an

activating effect. Therefore, cortical activation causes additional cortical ex-

citation through this striatal–thalamic loop. Among the structures involved

in the loop, the neostriatum is known to include many neurons that can in-

hibit their local neighbors and, by way of indirect connections, can have an

inhibitory or excitatory effect on more distant neostriatal neurons as well.It thus appears that if neurons in the neostriatum are being activated by

their cortical input, a complex pattern of selective inhibition and excitation

is being produced (Wickens, 1993).

In essence, there is a subcortical loop, including neostriatum, pallidum,

and thalamus, through which the cortex can stimulate itself. In addition, thereis the inhibitory network in the neostriatum that can be the basis of competi-

tion. This architecture has been proposed to realize a regulation mechanism

also affecting cortical activity dynamics (Wickens, 1993). One proposal wasthat distributed cortical neuron ensembles are linked to small populations of 

neurons in the neostriatum, paleostriatum, and thalamus, and that each cor-

tical neuronal assembly self-activates through the cortico–striato–thalamicloop. Given that each cortical assembly has this subcortical extension, the

striatal inhibitory neurons can provide the basis for competition between the

different ensembles. This is schematically illustrated in Figure 5.2, in which

open and filled circles are thought to represent two neuron ensembles. Note

that the two overlap in their cortical part, but inhibit each other by way of the inhibitory connections between their neurons in the neostriatum (Miller& Wickens, 1991; Wickens, 1993).

These considerations demonstrate that feedback regulation of cortical

activity can be implemented in the brain. Proposals for such implementation

are available and in agreement with cortical structureand function. The exact

wiring must be clarified by future research.

Page 98: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 98/331

5.1 Regulation of Cortical Activity 81

Figure 5.2. A possible realization of a regulation device in the brain is illustrated. Corticalneuron ensembles are assumed to be strongly connected to subcortical neurons in theneostiatum, paleostriatum, and thalamus. Each cortical neuronal assembly can activate itself through this subcortical loop. Inhibition between cortical neuronal assemblies is provided

indirectly by the inhibitory connections in the neostriatum. Filled and open circles indicatetwo overlapping webs. The subcortical inhibition can prevent simultaneous ignition of bothoverlapping webs.

5.1.6 Summary

The problem of cortical activity regulation can be minimized if correlation

learning, sparse coding, and feedback regulation are being applied. This

seems plausible neurobiologically. The basic idea therefore becomes that

correlation learning leads to the formation of sparsely coded representa-

tions of objects and words, each by one functional web, and that feedback

regulation keeps the cortical activity level within certain bounds.

The feedback regulation mechanism can be described in cybernetic terms,

and there are also proposals to spell it out in the language of neuronal

circuits. One possibly is that the inhibitory neurons in the neostriatum play

a role. Full activation of one of the neuron ensembles would, by way of 

striatal inhibition, prevent other competing networks becoming active as

well. In case of lack of strong activity in the cortex, the positive feedback loop

through subcortical nuclei could enhance the cortical activity level gradually.

The inhibition between neuronal assemblies can be provided even if the

cortical ensembles overlap – that is, if they share neurons. The illustrated

circuit shows that it is possible to keep overlapping cortical representations

and activity patterns functionally separate. It is therefore possible to realize

functionally discrete engrams by overlapping neuron ensembles.

Page 99: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 99/331

82 Regulation, Overlap, and Web Tails

5.2 Overlapping Representations

How would functional webs representing related words be organized? Asimilar question is, how can different readings of the same word or two

different ways to realize it phonologically, be modeled? The general an-

swer proposed for these questions will be that related items are organized

as overlapping cortical neuron ensembles. Putative mechanisms for suchoverlapping representations have been specified in Section 5.1. In general,

these mechanisms allow each neuronal representation α to become fully

active, whereas, at the same time, preventing those neurons of other over-

lapping ensembles β, γ , . . . , ω that are not included in α from becomingstrongly active as well. In the proposed jargon of functional webs, the more

precise formulation is the following. The ignition of α prevents other websβ, γ , . . . , ω, some of which overlap with α, from igniting at the same time.

The overlap areas α shares with β, γ , . . . , or ω would be activated as part of 

the ignition of α.

The following proposal is that overlapping but mutually exclusive dis-

tributed representations underlie the processing and representation of dif-

ferent types of related words. As pointed out in Chapter 4, functional websrepresenting individual words may include

   r a phonological or word-form related part located in perisylvian language

areas and strongly lateralized to the left   r a semantically related part distributed over various areas of both hemi-

spheres, whose topography may depend on semantic word properties.

The two parts, the form/phonological and semantic subwebs, would be func-

tionally linked, and thus exhibit (i) similar dynamics and (ii) mutual func-tional dependence.

In this framework, the representation of phonologically and semantically

related words is straightforward. Semantically related words should overlapin the semantic, widely distributed, bihemispheric part of their functional

webs, whereas phonologically related words should share neurons in the

perisylvian lateralized part of their functional webs.

The overlap between representations also assumed by many other psy-

cholinguistic theories (Section 6.2) explains why words closely related inmeaning and form can influence or prime each other. In this context, to prime means that, in experiments investigating word processing, the pre-

sentation of one of the related items, also called the prime, influences the

processing of the other related item that is presented later, usually called

the target . Priming effects can be facilitatory or inhibitory, that is, the prime

word can improve or deteriorate the processing of the target word. Both

Page 100: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 100/331

5.2 Overlapping Representations 83

semantically and form-related words can have a facilitatory priming effect

on each other (Humphreys, Evett, & Taylor, 1982; Meyer & Schvaneveldt,

1971). It must be mentioned, however, that the paradigm and, in particular,the onset with which the stimulus words are being presented can have an

effect on whether a priming effect is obtained, and on whether it is facilita-tory or inhibitory (Glaser & Dungelhoff, 1984; Levelt et al., 1991; Mohr &

Pulvermuller, 2002).

The issues of semantic and form-related overlap of functional webs is

discussed in greater depth in the following subsections.

5.2.1 Homophones and Form-Related Words

Syllables such as “spy” have at least two possible meanings with only minor,if any, relation to each other. How would such homophonous (or polysemic)

words be organized in a model of functional webs? The syllable may occur

frequently in the contexts of spiders and secret agents. The web represent-

ing the word form could, therefore, develop connections to the neuronal

counterparts of both word meanings. As argued in Chapter 4, these word-

form related or phonological webs should be localized in the perisylviancortex, whereas the meaning-related neurons would primarily be expected

outside the perisylvian cortex, in extra-perisylvian space. In addition, only

the form-related assemblies would exhibit strong laterality.

Now the problem arises that perception of the syllable “spy” will acti-

vate both semantic representations at a time. This is not desired; the twopossible word meanings should be kept separate. Conceptually, it is best to

speak about two word representations, each with its specific meaning, that

share their phonological representation. Likewise, it appears appropriate topostulate two neuron ensembles, one for each word, that overlap in their

perisylvian phonological part.

The mutual exclusion of the two overlapping functional webs could beprovided by a regulation mechanism of the type discussed in Section 5.1.

A scenario for this would be as follows: A word stimulus would activate

the phonological part shared by the two functional webs representing the

two words. Because the semantic parts of both word webs are strongly con-

nected to the shared phonological representation, they both receive exci-tation. There is a race between the two overlapping word webs that finallyends in the ignition of one of them, the one reaching its activation threshold

first. The outcome of the race is likely determined by the internal connec-

tion strength of each of the overlapping webs – the web of the more frequent

homophone being more likely to ignite first – and by activity in the webs due

to input in the past, for example, preactivation of the web of the word that

Page 101: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 101/331

84 Regulation, Overlap, and Web Tails

best fits into the context. The ignition of one of the word webs activates the

regulation mechanism, thus preventing the ignition of the competitor. Nev-

ertheless, as detailed, the competitor web initially profits from the activationprocess because it is partially activated in the initial phase of the activation

process, before the ignition takes place.

This scenario is reminiscent of psycholinguistic ideas developed on thebasis of priming studies (Swinney et al., 1979). Reaction time studies indi-

cated that presentation of a homophonous word form first leads to partial

activation of all of its homophone meanings, whereas only at a later point in

time is one of them being selected while the others are suppressed. Strongly

connected overlapping word webs that activate each other, but cannot igniteat the same time as a result of regulation processes, may be the mechanismunderlying this; initial broad partial activation of all overlapping webs and

later selective ignition of one of them would be the tentative neurobiological

realization.

This mechanism is not necessarily restricted to real homophones or poly-

semes. The same process could underlie the decision between different

“submeanings” or uses of one particular word. School , for example, can

either refer to a place (“the school in Brentwood”), an event in such a place(“school begins at nine”), a body of persons (“the school of Tesnierian lin-

guists”), or an animal group (“a school of fish”). These readings are related,

but nevertheless exclude each other in a particular context. Their putative

neuronal counterparts are overlapping functional webs that share some of 

their neurons in perisylvian areas, but inhibit each other by way of a regula-tion mechanism.

Homophones share their word form. There are also words composed

of other words, and similar arguments can be made with regard to them.In this case, the form of one word includes the form of another one. This

is called form inclusion and the included word form is distinguished fromthe including word form. This relation of form inclusion may hold for the

words’ phonology, orthography, or both. Examples of word pairs one of 

which includes the other are, for example, scan vs. scant , pair vs. repair , anddepart  vs. department . In these cases, the relationship between the words

concerns their formal aspects, but not their meaning. The representation of 

the form of the smaller item that is part of the larger word can be envisagedto be organized as sets of neurons, one of which includes the other. This idea

is sketched in Figure 5.3 in the diagram on the lower left.

It may not be appropriate to postulate that the form-related neuronal

populations fully include each other, because the pronunciation of the in-

cluded item is usually affected by the context of the larger one. There-

fore, the included item pronounced in other contexts can exhibit acoustic

Page 102: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 102/331

5.2 Overlapping Representations 85

Figure 5.3. Putative organization and cortical distribution of functional webs represent-ing related words. Each functional web is represented schematically as two ovals: one forits semantic part and the other for the phonological or word-form related part. The ovalswould therefore represent sets of phonological and semantic neurons. Overlap betweenovals means that webs share neurons. As discussed in Chapter 4, the semantic subassem-

bly would be expected to have most of its neurons located outside the perisylvian areas,whereas the phonological subassembly would be placed in the perisylvian areas. Lines rep-resent connection bundles linking neurons within a functional web. (Upper left) The dia-gram sketches the proposed organization of functional webs realizing homophonous words(“spy”), or different readings of the same word (“school”). (Upper right) Putative correlateof synonymous – or near synonymous – words (“car” and “automobile”). (Lower left) If oneword form includes another word form (“repair” and “pair”; this is dubbed form inclusion),the relation of inclusion may also hold for the respective phonological webs. (Lower right)In the case of  hyperonymy , if one term is more general than another (animal vs. dog), therelation of inclusion may hold for the semantic parts of the webs.

features it lacks in the context of the larger word that includes it. It may

thus appear more appropriate to postulate overlapping phonological repre-

sentations rather than that one is fully included in the other (as the writtenrepresentation would suggest). Still, however, there would not be too many

specific form features – and therefore neurons – of the small items, but

several specific features – and neuronal correlates thereof – of the larger

item.

It is particularly difficult for an autoassociative memory to keep separatetwo representations, one of which fully includes the other. A mechanism forthis could be envisaged, for example, on the basis of the striatal inhibition

model, as sketched in Section 5.1. However, it is certainly easier to sepa-

rate two distributed representations if each includes a substantial number

of neurons that distinguishes it from the other representation. Ideally, the

number of neurons shared between two ensembles should be small, certainly

Page 103: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 103/331

86 Regulation, Overlap, and Web Tails

not larger than the number of the neurons distinguishing one web from the

other. The larger the percentage of overlapping neurons becomes, the more

difficult it becomes to keep two representations apart.It would appear that the brain uses an appropriate strategy for overcom-

ing the overlap problem. Word forms that include others can be attached todifferent semantic representations. These semantic representations differ-

entiating between the similar forms can, so to speak, be used as handles to

keep the word forms separate. The non-overlapping parts, rather than sep-

arate neurons representing lexical items, can then be the target on which a

regulation mechanism can operate in order to exclude full activation of one

of the ensembles while the other one ignites. It is clear that various excita-tory and inhibitory processes could occur between such partially overlapping

networks of neurons if they are stimulated while activity is controlled by a

regulation mechanism. Nevertheless, the overlap would make facilitatory

effects possible.

Again, mutual exclusion of ignitions of overlapping neuronal representa-

tions in a regulated system profits from relatively small overlap regions. Suchmutual exclusion is difficult in a system of fully inclusive neuronal ensem-

bles. If form representations strongly overlap, it is possible to maximize thedistinguishing neurons by adding completely different semantic web parts

to the similar form representations. This may be so if word forms have very

different meanings. Many lexicalized or semantically opaque words include

at least one part of their form that can also be used in a completely dif-ferent meaning. The meaning change results, for example, if a derivational

affix is attached to a noun or verb stem, as, for example, in the previously

mentioned pairs repair  vs. pair  and department  vs. depart . In a model of 

functional webs, they would be realized as form-inclusive ensembles, assketched in the diagram on the lower left in Figure 5.3.

The representation is different for very similar forms whose meaning isclosely related and can be deduced from the meanings of the morphemes

they are composed of. These semantically transparent words, such as unpair 

or departure, can be conceptualized as being composed of their parts at

both the word form and the meaning level. The words depart  and depar-

ture have the same possible referent actions. Representations of words such

as departure would be composed of the two distinct representations of de- part  and the derivational affix (Marslen-Wilson, Tyler, Waksler, & Older,

1994). This proposal is consistent with the priming effects observed between

semantically transparent composites and their submorphemes (departure

and depart ), which appears to be more solid and stable than priming be-

tween semantically opaque forms and their elements (department vs. depart )(Marslen-Wilson et al., 1994; Rastle et al., 2000).

Page 104: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 104/331

5.2 Overlapping Representations 87

Translated into the language of functional webs, this may mean that a se-

mantically transparent complex form is organized as a widelydistributed web

(realizing a content word) and a more focused web (realizing the derivationalaffix). These two networks would be connected serially, by what is called a sequencing unit , or sequence detector in Chapters 9–12. Facilitation effects

wouldbetheresultoftheactivationofthesamenetworksbyprimeandtargetstimuli. Instead, the relation between a lexicalized or semantically opaque

form and its parts is more complex, as suggested by Figure 5.3, lower left

diagram. Some facilitation could, in this case, be the result of the overlap

of form representations, but the competing meaning representations would

work against it.

5.2.2 Synonyms

The mirror image, so to speak, of homophony is homonymy or synonymy.

Two different word forms share their meaning – or are very similar in mean-

ing. Examples would be automobile and car . The relation between synonyms

may be realized cortically by functional webs sharing, or largely overlapping

in, their semantic, mainly extraperisylvian, part.A decision between the two overlapping representations is necessary

when the semantic part shared by the neuron ensembles are activated. If 

a picture must be named, the overlap region of two synonyms (or near

synonyms) would be activated to the extent that some activity would also

spread to each of the word-form representations. The best activated word

web would ignite first, thus inhibiting the competitor through the regulationmechanism. Again, activity already present in the webs at the start and their

internal wiring strength may be crucial for deciding which of the alterna-tives is being selected. The average internal connection strength of the web

is likely influenced by the frequency with which the word is used in language.

The upper right diagram in Figure 5.3 gives an idea of how the repre-

sentation of synonyms may be realized in the brain; by identical or stronglyoverlapping widely distributed representations of the meaning and distinct

perisylvian left-lateralized subwebs for the different word forms. This is es-

sentially the mirror image of the homophony/polysemy diagram on the lower

left. Therefore, one may propose that the general activity dynamics and inter-ference effects should resemble each other. This is, however, not necessarilythe case because the overlap of homophone representations is necessarily

with respect to a temporally structured neuronal unit, whereas a possible

temporal structure in a semantic web part may be disputed (see Chapter 8).

A word can be realized in different ways. Different realizations of the

same word, for example, its pronunciations in different dialects, can be

Page 105: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 105/331

88 Regulation, Overlap, and Web Tails

conceptualized using a similar mechanism as proposed for synonyms. In

contrast to synonyms, a strong overlap should exist not only in the semantic

parts of the webs, but also in the phonological parts. Chapter 8 addresses theissues of alternative realizations of the same word form.

The meaning of words can be organized hierarchically. Words such asanimal , dog, and greyhound refer to wide and more narrow categories of 

objects, and therefore a model based on functional webs would conceptualize

theirmeaningsasneuronsetsthatincludeeachother.Thelargercategory–to

which the hyperonym (“animal”) refers – would include a smaller category –

referred to by the hyponym (“dog”). The semantic part of the web of the

hyperonym would therefore include the semantic part of the hyponym’s web.This is illustrated by the diagram on the lower left of Figure 5.3. Clearly, the

overlap of semantic representations should allow for facilitatory priming

effects, whereas the alternative word forms can be the basis of competition.

5.2.3 Prototypes and Family Resemblance

Theclaimthatamoregeneralterm’ssemanticwebistheunionofallsemantic

representations of its hyponyms is in need of refinement. The lower-levelcategories share features, and the respective semantic representations can

therefore be conceptualized as sets of semantic feature neurons that overlap.

The name of the higher-level category can be used to refer to all objects the

hyponyms can refer to. Therefore, the higher-level category name correlates

best with this overlap of the hyponyms’ semantic features. Given correlationlearning is crucial, the connections between the word form of the more

general term and its semantics relies primarily on the strong connections

between word form web and the semantic overlap neurons.Figure 5.4 schematically represents the shared semantic features by the

intersectionofthreeovals.Thevisualfeaturesofthemouthsoreyesoftypical

frequently encountered animals, such as dog, cat, and mouse, may be thoughtto be located here. If a less prototypical animal lacks some of these features,

or even most of them – think of the words octopus or jellyfish as examples –

its semantic representation would not link into this overlap region (dashed

oval in Fig. 5.4). It is therefore clear that priming between the hyperonym

and the less prototypical hyponym would be much reduced compared to thepriming expected between hyperonym and prototypical hyponyms.

If a picture of an object is being presented to a person and he or she is

asked to produce a name for it, the activation race between possible word

representations would again be determined by their internal connection

strengths. The outcome of the race depends on the correlation among the

relevant neurons and, thus, for example, on the frequency with which the

Page 106: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 106/331

5.2 Overlapping Representations 89

Figure 5.4. Putative organization of semantic relationships between concepts: overlap of prototypes and family ressemblance. Ovals represent semantic web parts of individualwords. Overlap means shared semantic features between word representations and sharedfeature neurons between webs. Think of the overlap area to include neurons responding tovisual features of mouth, eyes, and heads of typical animals. Three overlapping semantic webparts of prototypical concepts are illustrated by ovals in solid lines (e.g., cat , frog and shark ).A fourth semantic web representing a nonprototypical concept may not link into the over-lap area of the three prototypical webs, but nevertheless exhibits family ressembence with

them – that is, it would share some semantic feature neurons with some of the other webs(e.g., jellyfish). The union of all of these semantic representations may be considered thesemantic part of a web realizing a hyperonym of the words whose meanings are representedby the individual ovals (e.g., animal ). However, the sound-to-meaning correspondence of the hyperonym would be realized primarily as connections between the overlap area andits word-form representation.

respective words are used (their lexical frequency). Frequently occurring

basic-level category names, such as dog, may therefore be preferred to rare,

more specific terms, such as greyhound. A higher-level category name willlose the race against a more specific term (animal vs. dog) primarily because

of the better correlation between the specific set of activated semantically

related features on the one hand and the basic category name on the other.An additional factor is that higher-level categories can refer to very different

objects or actions, as Wittgenstein (1953) illustrated using the word game as

an example. The correlation among the semantic features these words relate

to is therefore necessarily low, resulting in relatively weak average links

within the semantic part of their webs. The word web of dog would thereforeignite faster than that of the more general term animal – because of its bettermeaning-form correlation and also faster than the web of the more specific

term greyhound, because of its higher lexical frequency. (Additional factors

may also be relevant.) These considerations are reminiscent of the priority

of basic level categories of the dog type in cognitive processing as made

evident by behavioral testing (Rosch et al., 1976).

Page 107: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 107/331

Page 108: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 108/331

5.3 Web Tails 91

5.3 Web Tails

This section discusses two very different issues that may nevertheless have asimilar neurobiological basis: How may the cortical representation of writtenwords link into the functional webs set up for words and what may be the

putative neurobiological correlates of affective meaning?

5.3.1 Affective and Emotional Meaning

The idea that cortical neuron webs may be linked to additional subcortical

circuits is discussed in Section 5.1. Evidence for the involvement of thalamicand striatal circuits in language processing comes from clinical studies. Fur-

ther, emotional and affective word properties may be represented by neu-

rons in the amygdala and midbrain. As a result of correlation learning, these

subcortical neurons may be attached as a tail, so to speak, to the cortical neu-ron ensembles representing words (Pulvermuller & Schumann, 1994). There

have been several tentative explanations of the variable success of learners

who started to learn their second language late in life. One proposal assumes

differential attachment of amygdala–midbrain tails to word webs. Only if word processing and subcortical emotionally related brain activity corre-

late, the learners’ word webs were envisaged to grow amygdala–midbraintails. Activation of these networks would flood the forebrain with dopamine,

thereby possibly facilitating further language-related learning. Therefore, a

person learning a second language who has already acquired a few word

webs with amygdala–midbrain tails may facilitate storage of more informa-

tion, including language and grammatically related information, when these

networks are active.Although the proposal has led to research efforts that yielded data con-

sistent with the basic idea at both the behavioral level (Schumann, 1997) and

that of neurophysiological investigation (Montoya et al., 1996), a systematic

investigation using functional brain imaging techniques must await future

research.

5.3.2 Linking Phonological, Orthographical,

and Meaning-Related Information

In Chapters 4 and 5, the visual representations of written words were largelyignored. The webs realizing word forms were proposed to be phonologically

related. Clearly, however, literate speakers must have acquired learned rep-

resentations of written words and their neurobiological counterparts are

likely localized in visual areas of the occipital and inferior temporal lobes.

Equally clearly, however, the written word-form representations should be

Page 109: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 109/331

92 Regulation, Overlap, and Web Tails

Ellis and Young, 1988

Connections proposed 

From ↓ to→

   A  u   d   i   t  o  r  y  a  n  a   l  y  s   i  s   l  e  v  e   l

   P   h  o  n  e  m  e  o  u   t  p  u   t   l  e  v  e   l

   A  u   d   i   t  o  r  y   i  n  p  u   t   l  e  x   i  c

  o  n

   S  p  e  e  c   h  o  u   t  p  u   t   l  e  x   i  c  o  n

   S  e  m  a  n   t   i  c  s  y  s   t  e  m

   V   i  s  u  a   l   i  n  p  u   t   l  e  x   i  c  o  n

   G  r  a  p   h  e  m   i  c  o  u   t  p  u   t   l  e

  x   i  c  o  n

   V   i  s  u  a   l  a  n  a   l  y  s   i  s  s  y  s   t  e  m

   G  r  a  p   h  e  m  e  o  u   t  p  u   t   l  e  v  e   l

Auditory analysis level

Phoneme output levelAuditory input lexicon

Speech output lexicon

Semantic system

Visual input lexicon

Graphemic output lexicon

Visual analysis system

Grapheme output level

Autoassociative

 Network model

Connections proposed 

From neurons representing ↓

To neurons representing→

   A  c  o  u  s   t   i  c   f  e  a   t  u  r  e  s

  o   f  s  o  u  n   d  s  -   i  n  -  c  o  n   t  e  x   t

   A  r   t   i  c  u   l  a   t  o  r  y   f  e  a   t  u  r  e  o   f  s  o  u  n   d  s  -   i  n  -  c  o  n   t  e  x   t

   A  c  o  u  s   t   i  c  w  o  r   d   f  o  r

  m

   A  r   t   i  c  u   l  a   t  o  r  y  w  o  r   d

   f  o  r  m

   W  o  r   d  m  e  a  n   i  n  g

   V   i  s  u  a   l  w  o  r   d   f  o  r  m

   W  r   i   t   i  n  g  p  a   t   t  e  r  n  o   f

  w  o  r   d   f  o  r  m

   V   i  s  u  a   l   f  e  a   t  u  r  e  s  o   f

   l  e   t   t  e  r  s  -   i  n  -  c  o  n   t  e  x   t

   F  e  a   t  u  r  e  s  o   f  w  r   i   t   i  n  g  g  e  s   t  u  r  e  s  -   i  n  -  c  o  n   t  e  x   t

Acoustic features of sounds-in-context

Articulatory feature of sounds-in-context

Acoustic word formArticulatory word form

Word meaning

Visual word form

Writing pattern of word form

Visual features of letters-in-context

Features of writing gestures-in-context

Page 110: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 110/331

5.3 Web Tails 93

coupled to the perisylvian phonological webs, and this should be so for all

word categories alike. Therefore, no predictions on word-category differ-

ences can be derived, except if the correspondence of written and spokenword form differs, as is the case for words with regular and irregular spelling

(home vs. come).

Correlations are present at the levels of features of letters and phonemesand at the level of spoken and written word forms. Obviously, letter–sound

correlation plays a greater role for regularly spelled words than for irregu-

lar ones. Additional correlation exists between aspects of the written word

form and aspects of referential meaning. The correlation between individual

letters and meaning aspects might not be high in most cases; however, if let-ters and their local context are assumed to be processed as units below theword and morpheme level, relevant correlation between letters-in-context

and meaning aspects can also be postulated. The proposal is that neurons

related to context-sensitive variants of a phoneme or grapheme would link

with semantic neurons. This suggestion is worked out in greater detail with

regard to phonemes in Chapter 8.

Neuropsychological models in the tradition of Morton’s logogen model

(1969) have shed light on the complex relationship between systems involvedin the processing of input- and output-related information about spoken

and written words. Autonomous processors, modules for the production and

perception of spoken and written words, were postulated and related to

modality-specific lexical modules and one central semantic system (Ellis

& Young, 1988).The assumption of a uniform central semantic system contrasts with other

approaches in which much emphasis is put on multiple semantic systems pro-

cessing different types of category-specific semantic information. Chapter 4summarized such an approach.

Figure 5.5. (Upper matrix) Matrix display of the core part of the modular model of wordprocessing proposed by Ellis and Young (1988). The terms in the rows and columns refer toindividual modules. Each dot represents a connection from a module indicated in the leftcolumn to a module indicated in the top line. Filled dots refer to within-module connections.Open dots refer to connections between modules. Only 29% of the possible between-module connections are realized. (Lower matrix) If information about spoken and written

word forms and about meaning is thought to be processed by neurons in an autoassociativememory, all possible connections should be present and effective. In this diagram, proposedlocal cortical connections are indicated by filled dots and long-distance cortical connectionsare indicated by open dots. The connections can be taken to define the internal structureof one word web. The matrix lacks detail about important aspects of the neurobiologicalmodel. For example, it does not specify category-specific semantic systems and glossesover the fact that phonologically related neurons may be included in (rather than connectedto) the web realizing the word form. Nonetheless, the matrices stress a potentially importantdifference between a modular approach and an autoassociative memory model.

Page 111: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 111/331

94 Regulation, Overlap, and Web Tails

The main proposal of modular models of word processing is that the men-

tioned types of word-related information are processed in separate modules

with specific selective connections between them. Figure 5.5 illustrates themodules and connections postulated by one of the most developed modular

models of word processing (Ellis & Young, 1988). The model is presented

here in the form of an autoassociation matrix in which dot signs representconnections and blank squares represent their absence. Each term appearing

on the left (and again appearing at the top) refers to a distinct functionally

autonomous system, a module. It can be seen that more than half of the

81 possible connections between modules are assumed to be not present. It

appears appropriate to assume very strong connections within each of thenine modules, because information processing is primarily taking place there.These within-module connections are indicated by the black dots on the di-

agonal. Between modules, only a minority of the possible connection is pro-

posed to be realized (21 out of 72 possible ones, which equals 29 percent).

The sparseness of connections is motivated by the research strategy applied.

The zero assumption was no connection, and each connection was required

to be motivated by a double dissociation or by at least one patient whose

behavior was interpretable as evidence that the respective type of informa-tion transfer was specifically impaired. Thus, insertion of an open dot (or

arrow in the original diagrams) needed justification by neuropsychological

dissociations.

An associative memory model of the respective types of information ex-change would suggest a different conceptual approach. The cortex would be

considered an autoassociative memory in which correlation causes strong

links. It cannot be disputed that writing and seeing a letter is accompanied

by neuronal activity in visual and motor cortical areas, including primary butprobably extending into higher-order areas. Because as a rule the frontal and

temporo-occipital systems are strongly linked and connections are recipro-cal, it makes sense to postulate that strong reciprocal connections develop

between the visual and graphic representations. This leads to the proposal

that additional reciprocal connections between writing gestures and visual

analysis (not present in the original matrix) should exist. Furthermore, young

children learning to write would usually articulate and/or hear the respective

phonemes/graphemes. Thus, reciprocal connections between auditory andvisual phoneme and grapheme systems should develop as well. This leads to

the addition of four more connections, linking the two analysis systems to

each other and to the output systems also. The same argument should hold

at the level of word forms, and the semantic system should have reciprocal

connections to all other systems. A difference between models would be

that the cotext-sensitive variants of phonemes, and possibly also graphemes,

Page 112: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 112/331

5.4 Summary 95

would be thought to be part of (rather than connected to) the word-form

representations in the respective lexicons.

In summary, the postulate put forward here is that strong connections arepresent between all components of a distributed representation of a word.The zero assumption would therefore be that neurons involved in the pro-

cessing of one aspect of a word are linked directly to neurons involved in

the processing of all its other aspects. Neuropsychological dissociations can

still be explained within the autoassociative approach, because cutting one

particular connection – together with the partial lesion of adjacent represen-

tations – may still cause a processing defict primarily affecting performance

on specific tasks (see also Excursus I).In conclusion, a neurobiological model can incorporate aspects of visual

word processing. The neuronal webs may include attachments or tails that

are connected reciprocally to the neuron sets representing the phonological

word form and the word’s semantics. It must be left for future research

to show how an autoassociative network model can account for the greatvariety of neuropsychological symptom constellations involving reading and

writing.

5.4 Summary

This chapter aimed at refining aspects of the functional web model of lan-

guage processing outlined in Chapter 4. Mechanisms for maintaining the

cortical equilibrium of activity were discussed in Section 5.1, and complex

sound-to-meaning relationships of words were treated in Section 5.2. Finally,

the possibility was mentioned that word webs may have additional neuron

populations attached to them that could relate to knowledge about the affec-tive valence of the words and to the written image of a word and respective

actions involved in writing it.

The regulation mechanism discussed in Section 5.1 plays an important role

in the grammar mechanisms discussed in Chapters 10 to 12. The regulation

mechanism establishing inhibition between strongly connected neuronal en-

sembles can be considered a mechanism that explains a universal feature of language occuring at various levels. It could be important for the mutual

exclusion of two word forms with the same meaning, of two pronunciationsof the same word, of two meanings of homophonous words, of two readings

of the same word, and even of two interpretations of the same sentence

(word string).

Page 113: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 113/331

CHAPTER SIX

Neural Algorithms and Neural Networks

What does the investigation of artificial models of networks of neurons con-

tribute to the investigation of brain function in general, and of language

mechanisms in particular? There are at least two possible answers to this

question.

One answer is that an artificial neural model can be used to prove that

a circuit of a certain type can solve a given problem. When thinking aboutcomplex interactions of functional neuronal units, one is in danger of losing

track of what a network actually can, and cannot, do. Here, a simulation may

help by providing an existence proof that a desired result can be obtained

using a given circuitry or circuit structure. A successful simulation never

proves that the network it is based on is actually realized in the nervous

system; it shows only that the network type the simulation is based on is

one candidate. To call it a realistic simulation, other criteria must be met:

in particular, that the network structure can be likened to brain structureand that the functional principles governing the neurons’ behavior and their

actual behavior have analogs in reality as well.

A second possible answer to the question about the significance of neu-

ron models is that they can serve as illustrations of one’s ideas about brain

function. In the same way as a detailed verbal description or a description in

terms of algorithms, a neuron circuit can help to make an idea more plastic.

However, algorithms or verbal descriptions of mechanisms cannot always

be likened to processing elements in the brain. In contrast, model circuitsalways have some relationship to potentially real neuronal mechanisms, and

it is possible to make such a model more realistic by including more detail

about neuroanatomy and neurophysiology into the network’s structure and

functioning.

Historically, neural networks have been of utmost importance in a dis-

cipline one may want to call brain theory, the systematic theoretical

96

Page 114: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 114/331

6.1 McCulloch and Pitts’s Logical Calculus as a Starting Point 97

investigation of the function of the brain. Groundbreaking work started

in the 1940s with McCulloch and Pitts’ work, in which symbols called arti-

  ficial neuronswere used to illustrate ideas about how neurons may realizelogical operations. This chapter introduces these ideas, gives a brief overview

of some neural network architectures, and highlights a few applications tolanguage-related problems.

Some of the network architectures featured in this chapter exhibit fam-

ily resemblance to the cell assembly framework outlined in earlier chapters.

One obvious difference between the neural network approaches to language

discussed later and the cell assembly framework is that the latter provides

explicit statements about the brain areas in which neuronal processes re-lated to cognition are postulated. Such statements are sometimes lacking in

neural networks frameworks in which cognitive processes are assigned to

artificial neurons ordered in layers or modules but not to brain areas. It is,

however, not a principle problem to relate neuron layers to brain structures.

Differences and common features between different types of neural network

proposals are discussed in more detail.

6.1 McCulloch and Pitts’s Logical Calculus as a Starting Point

In the early 1940s, McCulloch and Pitts published a paper titled “A Logi-

cal Calculus of Ideas Immanent in Nervous Activity” (1943). In this article,

the authors discuss putative brain mechanisms possibly underlying the abil-

ity to recognize complex sequences of events. Their “calculus” was laterreformulated and further developed as a theory of finite-state automata

(Kleene, 1956; Minsky, 1972). Considerations were based on circuits made

up of simple computational elements that share properties with nerve cells.The question was to which type of events such artificial neurons could re-

spond to specifically. In this context, the term event refers either to stimuli

and stimulus constellations in the environment or to neural events – that is,activity patterns of other artificial neurons spread out in space and time. The

computational units would therefore specifically respond to, or “recognize,”

spatiotemporal patterns of activity (and stimuli), and analogous circuits can

be constructed as producing, or “generating,” similarly complex events.

The proposed artificial computational elements were later calledMcCulloch–Pitts-neurons. They transform an input into an output. The out-put is binary, and the neuron is either active or inactive. This binary character

is inspired by the knowledge about real neurons that, at each point in time,

either generate an action potential as an output or not. Similar to their real

counterparts, the artificial neurons are active at a certain point in time if they

have received enough excitatory input one moment, or time step, before. For

Page 115: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 115/331

98 Neural Algorithms and Neural Networks

Table 6.1. The logical operations carried out by the circuits in Figure. 6.1are OR, AND, NOT, and EITHER–OR (or XOR). The truth conditions of

these operations are summarized. Variables a and b can be thought of asstatements that can be either true (T) or false (F). The table lists theresulting truth values of statements that include the operators. If Tis replaced by 1 and F by 0, and a and b by the neuron labels α and β,the activity value of the rightmost “cardinal” cell in the four diagramsof Fig. 6.1 can be derived from the activity states of the leftmost“input neurons.”

a b a OR b a AND b NOT a a XOR b

T T T T F F

T F T F F T

F T T F T T

F F F F T F

ease of illustration, the time continuum is, so to speak, sliced into discrete

time steps.

The artificial neurons vary with regard to the number of simultaneousinputs they need for becoming active. One neuron may be activated by only

one input; it would have a threshold of 1, whereas others may have a higher

threshold, for instance, of 2, 3, or 10, and would therefore need at least

2, 3, or 10 simultaneous excitatory inputs to become active. The artificial

neurons are usually symbolized by circles, which can be thought to representthe analog of cell body plus dendrites of real neurons, and by arrows pointing

to other neurons that may be considered to be analogs of axons and their

branches.The neuron circuits one can build using these artificial threshold neu-

rons can realize logical operations (see Table 6.1). Two neurons α and β

may project to a third neuron γ , whose threshold is 1 (Fig. 6.1, uppermostdiagram). This circuit can be considered a representation of a logical “Or”

operation. The third neuron, γ , becomes active under three conditions: if α

is active, if β is active, or if both α and β are active at the same time step. If 

any of this happens at time t , the neuron γ  fires at the next time step, t + 1.

Otherwise it remains silent, given it does not receive additional input fromother neurons. Activity of the neuron γ  would therefore signify the eventthat α or β were active just one time step earlier. One can say that the circuit

realizes, and the neuron γ  represents, the logical “Or” operation.

The neurons α or β may be sensory neurons that become active when spe-

cific stimuli are present in the environment. The stimuli, or stimulus features,

may be labeled a and b, respectively. In this case, the activity of the γ  neuron

Page 116: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 116/331

6.1 McCulloch and Pitts’s Logical Calculus as a Starting Point 99

Figure 6.1. Logical circuits. In each diagram, theright-most cardinal neuron specifically responds toan activity constellation of the input neurons on theleft. The cardinal neuron becomes active if neuronα OR neuron β is active (uppermost circuit), if neu-rons α AND β are simultaneously active (secondfrom top), if neuron α was NOT active (third from

top), and if EITHER neuron α OR neuron β havebeen active (“XOR” circuit at the bottom). Circlesdenote artificial neurons and arrows directed con-nections of strength 1. The activation threshold isindicated by a number in the circle. (For furtherexplanation, see text.)

would represent the complex event that a or b was present in the input.

The event c indicated by the γ  neuron’s activity would be equal to “a or b.”

To choose a more concrete example, the sensory neurons may respond tostimuli appearing at the same location of the visual field. One of them could

be responsive to green color and the other one to circle shapes. In this case,

the firing of the γ  neuron would tell one that an object with the feature of greenness or roundness was present. Therefore, it would fire if the object

was a green ball, an orange, or a crocodile.

By changing the threshold of the γ  neuron to 2, the function of the cir-cuit can be altered (Fig. 6.1, second diagram from top). In this case, two

simultaneous inputs, from neurons α and β, respectively, are necessary to

activate γ . This neuron would now only respond after both sensory neurons

had been active at the same time. Thus, the stimulus features a and b must

both be present. The neuron γ  would respond in the event of the perceptionof a green ball, but not to an orange or crocodile. This circuit now realizes alogical “And” operation.

The logical “Not” operation can also be translated into a simple network.

For this to work out, one may want to assume a neuron that is spontaneously

active – that is, a neuron that fires even if not stimulated before. As an

alternative, the neuron may receive continuous input from somewhere else.

Page 117: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 117/331

100 Neural Algorithms and Neural Networks

This neuron, γ , could now receive an inhibitory input from a sensory neuronα so that it would stop firing one time step after α was active. Activity of γ 

would therefore indicate thatα

was not active and, thus, that the stimulus towhich α responds was not present in the input.

More complex circuits of artificial McCulloch–Pitts neurons are necessary

for realizing other logical operations. One may wish to have a neuron thatresponds to green objects and to round objects, but not to objects that are

green and round at the same time. For this purpose, a circuit must be created

that realizes the “Exclusive Or” operation. This is not possible based on one

single McCulloch–Pitts neuron, but would require an arrangement of such

units in two steps, or layers.One possibility, also illustrated in the diagram at the bottom of Figure 6.1,

is the following: The input neurons α and β each activate one additional neu-

ron, γ  and δ, respectively. In addition, each of them inhibits the neuron the

other input neuron activates. Therefore, α activates γ  and inhibits δ, and β

inhibits γ  and activates δ. A fifth neuron, ε, receives activating input from

both γ and δ. All thresholds are 1, so that one activating input alone – withoutpresence of additional inhibition – suffices for firing any of the neurons. The

ε neuron now is activated if α was active while β was silent (because γ  fired)and if α was silent while β was active (because δ fired). Simultaneous exci-

tation of α and β does not effect this circuit because the inhibition caused

at the second processing step, the neurons in the middle layer, cancels the

effect of any activation. The circuit can be considered a neural analog of an“Either–Or” operation. Activation of the network’s cardinal cell, ε, repre-

sents the fact that either input feature a or input feature b was present in

the input. It would respond to the orange or to the crocodile, but not to the

green ball.Again, this example illustrates that for certain problems, in this case, the

calculation of an “Either–Or” operation, the solution cannot be achievedby one single neuron. It is necessary to introduce processing steps into the

network and arrange the computational elements in layers through which

activity flows serially. This is an important insight that dates back to the 1940s

and 1950s (Kleene, 1956; McCulloch & Pitts, 1943) and was critical for the

emergence of the strong interest of neural modelers in multilayer networks

in the 1980s and 1990s. A neuron arrangement with an input layer and twosubsequent processing steps can realize any event that can be described by

a logical formula including the logical operators “And,” “Or,” and “Not”

(Kleene, 1956).

Networks of McCulloch–Pitts neurons can not only be used to detect

simultaneous stimuli or events, they can be used to represent complex events

spread out in time as well. For example, the sequence of words in a simple

Page 118: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 118/331

6.1 McCulloch and Pitts’s Logical Calculus as a Starting Point 101

Figure 6.2. Illustration of the representation of a com-plex event – the word sequence “Betty get up” – in aMcCulloch–Pitts network. The neurons in the upper lineare called input units, each of which is activated by the oc-currence of a given word in the input. The neuron at thebottom is a string detector  that becomes active only if the word sequence occurs. Numbers indicate the thresh-old of neurons – that is, how many simultaneous inputsthey need to become to active. The string detector can

be considered a logical unit that indicates that the con- junction of three assertions is correct.

sentence can be represented by a simple network. Assuming neurons that

respond specifically to individual words of a language, the McCulloch–Pittsnetwork in Figure 6.2 can be used for detecting the word sequence (1):

(1) Betty get up.

The network shown in Figure 6.2 includes neurons, depicted at the top,that specifically respond to one of the three words. Because they respond

to a specific input, they can be called input units. At the very bottom of the

graph is a neuron that fires if the string of words had been presented and theinput units have therefore been fired in a given order . Because this neuron

at the bottom fires in response to a string of elementary events, one can call

it a string detector . Apart from the input units and the string detector, thenetwork is composed of neurons whose function is to delay activity caused

by the early events so that all relevant information arrives simultaneously

at the string detector. Because their function is to delay activation, they

can be called delay units. This type of network transforms a series of events

spread out in time, the serial activation of input units, into simultaneousevents, the simultaneous activation of the neurons feeding into the stringdetector. The string detector in Figure 6.2 has a threshold of 3, thus needing

three simultaneous inputs to become active. Therefore, the entire series of 

three words must be presented to the network to activate the string detector.

Missing a word, introducing a delay between words, or changing the order

of the words will not allow the string detector to become active.

Page 119: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 119/331

102 Neural Algorithms and Neural Networks

Figure 6.3. Modified McCulloch–Pitts network forsequence detection. Now, sequential activation of thesequence detectors at the bottom would be the criterionfor string acceptance.

In these example networks, the criterion for the detection of a complex

event by the network was always the activation of one single neuron. This

neuron, the string detector in Figure 6.2 and the neurons representing logi-

cal operations in Figure 6.1, can be called the cardinal cells of the respective

networks. Instead of having a cardinal cell in the network, it is also pos-

sible to introduce a criterion for event detection based on an activity pat-tern in which more than one neuron participates. Such a criterion could be

that input causes a continuous sequence of neuron firings. The network inFigure 6.3 could then serve the purpose of string detection, because the neu-

rons in the lower row only become active in sequence if one particular input

sequence occurs. In this case, several neurons would be involved in the stringdetection process (bottom row of neurons in Fig. 6.3), apart from the several

input units (top row of neurons in Figure 6.3). Delay neurons are no longer

necessary, because direct connections between the neurons and activation

thresholds include the information about the correct sequence. If string (1),

the string “Betty get up,” occurs in the input, this causes sequential activa-tion of the three units involved in string detection. If the same words arepresented in a different order, at least one of the string-detecting neurons

fails to become active.

One may consider most of the networks discussed so far as heterogeneous,

because their neurons have different thresholds. Activation thresholds of 

cortical pyramidal neurons do not differ so much; it is rather the connection

strength of synapses that differs widely between neuronal connections. One

maythereforearguethatthesenetworksareinneedofimprovementinorderto make them more realistic. This could be done by changing all thresholds

to 1 and varying connection strengths instead. Because connection strengths

between neurons are known to vary, varying connection strengths in artifi-

cial networks may be considered more realistic than varying their thresholds.

Note that in all diagrams, all connections were either excitatory (lines ending

Page 120: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 120/331

6.1 McCulloch and Pitts’s Logical Calculus as a Starting Point 103

in an arrows) or inhibitory (lines ending in a T shape). If, instead, connec-

tion strength was varied, this could be indicated by writing the strength or

weight of each link as a number next to the arrow or line. The “Or” circuit inFigure 6.1 would not need to be changed, in this case, the connections all re-ceiving the weight 1. However, the connections in the “And” network would

need to be changed to 0.5 so that the cardinal neuron, whose threshold would

be set to 1, would need two simultaneous inputs to become active. In the

same way, some of the connections in Figure 6.3 would need to be readjusted

to 0.5. The function of the circuits would not change as a consequence of 

these adjustments. The less realistic circuits depicted in the figures can nev-

ertheless be preferred for illustration purposes. One reason for preferringthem might be that they require specification of a smaller number of param-

eters and may, therefore, be considered easier to oversee. Note that in the

depicted networks, the number of connections that must be labeled is usu-

ally greater than the number of neurons whose threshold must be specified.

Depicting the less realistic circuits is also not a problem here because themore realistic equivalent solutions can be derived easily.

Figures 6.2 and 6.3 show two ways in which a sequences of events can be

mapped onto neuronal dynamics. In one case, a string detector is presentthat receives simultaneous input informing it about a specific serial activa-

tion pattern. In the second case, the sequence of stimuli is imitated by a

sequence of activation of neurons involved in the sequence detection pro-cess. As pointed out in Chapters 8 and 9, the two neuronal strategies of 

string detection that can be postulated on theoretical grounds appear to

have analogs in real neuron circuits.

The networks discussed so far can only detect events that occur within a

short time interval, usually a few time steps. The temporal extend of theseevents is certainly always finite. However, it was one of the main points of 

McCulloch and Pitts’s proposal that their networks can also detect events

of arbitrary length, endless strings in principle. This can be achieved by

introducing loops in the network.

Figure 6.4 presents two neurons that respond to strings of indefinite

length. On the left is a neuronal element that becomes active only if it wasactive at an initial time step, and if it received continuous input since its

initial activation. Thus, it is active later only if it was supplied with a contin-uous sequence of excitatory inputs. As soon as an input fails to stimulate the

neuron, its activity immediately ceases, and it is silent until reactivated by

a mechanism not specified here. Assuming that the neuron responds to the

input “very,” it is active if an arbitrarily long word chain including only thisparticular word has been perceived. One may consider this a somewhat odd

Page 121: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 121/331

104 Neural Algorithms and Neural Networks

Figure 6.4. McCulloch–Pitts neurons with self-connectingloops. (Left) The neuron with threshold 2 stays active if itwas active initially and received continuous input since. Its

activity means that a given input was always present (be-ginning at a predefined starting point). It can represent astring of indefinite length in which a symbol occurs repeat-edly. (Right) Activity of the neuron with threshold 1 indi-cates that a given input has occurred at least once. Thelatter neuron may be helpful for modeling discontin-uous constituents – that is, two morphemes that belongtogether although they are separated in a sentence.

device, but if the neuron is allowed to respond not to only one particularword, but instead to a class of words, for example, adjectives, its potential

usefulness for syntactic processing may become obvious. It could then detectlong strings of adjectives as they can occur preceding a noun (“the young,

beautiful, attractive, little, . . . frog”). Because the neuron stays active only

during a certain interval if a given input – for example, the word “very” –

is continuously present throughout all time steps, one may call it an “All”

neuron.

In contrast, the neuronal element on the right in Figure 6.4 becomesand stays active whenever the element it responds to has occurred once.

Because it continuously stimulates itself once it becomes active, the neuron

stays active for, in principle, an unlimited time. Whereas the “All” neuron’s

activity signals that a particular elementary event has occurred at all time

steps (during a certain interval), this present neuron’s excitation would mean

that at least one time step exists when a particular event was present. It couldthereforebecalledan “Existence”neuron.Neuronsofthesetypes,expressing

the existence and all operations, are necessary to neurally implement aspectsof the logical predicate calculus.

If neurons are so helpful in implementing logical calculi, why should they

not be helpful in modeling sentences? An “Existence” neuron could be

useful for storing the information that a particular word has occurred in thepast. The verb get and the particle up in sentence (1) belong together. When get has occurred, it would be advantageous to store this for some time, not

only because there may be a delay between the two words as a result of 

hesitation or breathing but also because other word material may intervenebetween the verb and its particle. An “Existence” neuron could store the

information that the verb has occurred and, via a direct connection, couldprime the particle’s neural representation so that its activation would be

facilitated later.

It is worthwhile to take a closer look at the possibilities opened by the

assumption of neuronal units that can, in principle, stay active forever.

Page 122: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 122/331

Page 123: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 123/331

106 Neural Algorithms and Neural Networks

the verb, with, in principle, no upper limit for the time delay between the

two.1

This example of the representation of two linguistic elements that belongtogether but are nevertheless separated in time or space shows that net-

works of McCulloch–Pitts neurons can solve linguistic problems that cannotbe solved by probabilistic grammars such as Markovian sources (Charniak,

1993; Markov, 1913). Markov chains and other so-called k-limited stochas-

tic models use the probabilities with which a (k+ 1)th element follows a

string of k other elements (obtained from large text corpora) to calculate

the probability with which a particular word (or letter) follows a given sen-

tence fragment.An example how this works is the following: Markov (1913) used Pushkin’s

novelEugeneOnegin to calculate the probabilities of vowels and consonants

to follow each other. He also calculated the probabilities of consonants (C)

and vowels (V) to follow series of CC, CV, VC, and VV. In this case, k = 2,

that is, only two preceding elements were used to calculate the conditional

probability of the next letter type. The same can be done for words in sen-tences and for largerk’ s (for further discussion, see Miller & Chomsky, 1963;

Shannon & Weaver, 1949).However, the fact that an element occurring in a string (e.g., the root

“turn”) leads to a 100 percent probability of occurrence of another element

(the verb particle “on”) at an indefinite time step later cannot be modeled in

the Markovian framework. To model it in the Markovian framework, the twodistant elements that belong together would have to be treated differently

from the intervening string. Both sentence (6), which includes a verb particle,

and sentence (7), which includes an intransitive verb without a particle.

(6) Betty steht morgens um halb drei Uhr morgens mit einem m ¨ urrischen

Gesicht auf . (including the particle)

(7) Betty geht morgens um halb drei Uhr morgens mit einem m ¨ urrischen

Gesicht. (without particle)

Nevertheless, the ten words preceding the particle in sentence (6) are the

same as the ten words at the end of sentence (7). Therefore, the differ-

ence between the two would escape a k-limited stochastic source if k = 10.

The probability for the particle to occur is primarily influenced by the dis-tant morpheme (with no strict upper limit for the distance) – not by its

1 This preliminary solution raises additional problems. First, the criterion for string acceptance

must be modified. Second, the bistable element must be switched off after the particle occurs

in the input. In later chapters, in particular 9 and 10, related ideas will be looked at in more

detail.

Page 124: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 124/331

6.2 Symbolic Connectionist Models of Language 107

next neighbors. To model this, probabilities would have to be calculated

for strings of arbitrary length. An extension of the Marcovian approach

is necessary – for example, so-called Hidden Marcov models – to capturethis property of sentences. A McCulloch–Pitts network seems to be able to

solve one serial-order problem in language that stochastic models of a par-ticular type have systematic difficulties with. One may ask in which way a

McCulloch–Pitts-based approached could help in capturing serial-order re-

lationships in language, where its principle limitations are, and how these

can be overcome.

McCulloch–Pitts’s framework allows for modeling logical operations at

the level of neurons. It can also be used to model serial order of complexspatiotemporal events. Historically, McCulloch and Pitts’s proposals were

crucial for the emergence of automata theory, without which the develop-

ment of modern computers would have been unthinkable. In language sci-

ence, they inspired research on grammar algorithms. Grammar models such

as finite state grammars and augmented transition networks were built on

the proposed neural mechanisms. These approaches to serial-order mecha-nisms were successful in modeling particularities of grammatical rules (see,

for example, Winograd, 1983), although the need to incorporate more so-phisticated rules sometimes led to extensions that made the resulting devices

somewhat more difficult to relate to neuronal structure. Examples are the

recursive embedding of networks and the addition of registers that are acted

on under special conditions whose implementation in terms of neurons mayappear opaque. This is not to say that these entities and processes could

not themselves be transformed into circuits of artificial neurons. However,

in the present context, the task is to provide candidate circuits for syntactic

processing that may be realized in the brain. The circuits originally proposedby McCulloch and Pitts seem to be closer to this general aim than some of 

the more recent and more sophisticated syntactic algorithms rooted in theirwork.

6.2 Symbolic Connectionist Models of Language

The idea to model language and cognitive processes on the basis of artificial

neuron-like devices was applied at different levels. Syntactic models, such asfinite state grammars, were proposed that modeled aspects of serial-ordermechanisms, and at the level of linguistic units – words and language sounds,

for example – processing models were proposed that specified the activation

processes envisaged to occur during production and comprehension.

According to modular approaches to cognitive psychology, the mental

language processor consists of quasi-autonomous subprocessors or modules,

Page 125: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 125/331

108 Neural Algorithms and Neural Networks

and therefore language processing is being considered the result of sub-

processes carried out by these. The subprocesses envisaged to be involved

in language comprehension are, for example, input-feature analysis, letteror phoneme analysis, word-form processing, and semantic analysis (Ellis

& Young, 1988). A similar but reverse cascade has been assumed for the pu-

tative subprocesses of language production that finally results in movementsof the articulators or the (writing) hand (Garrett, 1980). The postulated sub-

processes of word comprehension and production are assumed to occur se-

quentially or in an overlapping cascaded manner. It cannot be disputed that

during certain types of language production and comprehension, informa-

tion about stimulus features, letters or phonemes, word forms and semanticsare being processed. One may therefore refer to different levels or stages of language processing, such as the word form or lexical stage or the semantic

level. A specific proposal immanent to one class of modular models now is

that information processing is autonomous at each of the different levels.

In symbolic connectionist models, the subprocessors of modular models

have been replaced by layers of neuron-like elements, the assumption being

that individual artificial neurons or nodes represent acoustic or visual fea-

tures, phonemes or graphemes, word forms, and word meanings (Dell, 1986;Dell et al., 1997; McClelland & Rumelhart, 1981). Figure 6.6 presents one

model (Dell, 1986) in which each representation of a word is composed of a

central lexical node and its associated meaning-related (semantic) and form-

related (phonological) nodes. In this network, a stimulus word first activates

the phonological nodes corresponding to its sounds. These would, in turn,strongly activate the lexical node of the respective word and partially acti-

vate lexical nodes representing words phonologically similar to the stimulus

word. Activity then spreads to the semantic layer. This model has recip-rocal connections so that backward flow from the semantic to the lexical

layer and from the lexical to the phonological layer would also be allowed.

Computations at the different levels of the network can take place at thesame time and there can be a continuous exchange of information between

levels during computation. Each neuron of the network would at each time

step compute the sum of its inputs and receive an activity value accordingly.

In contrast to McCulloch–Pitts neurons, the networks’ neurons would not

need to have a fixed activation threshold, but their activity level may varycontinuously. A special criterion, for example, which node is most activein the lexical layer, may serve as the criterion for considering the word it

represents to be selected among alternative words whose nodes have also

been activated to some degree. The nodes are assumed to stay active for

some time after their stimulation, not only because they continuously send

Page 126: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 126/331

Page 127: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 127/331

110 Neural Algorithms and Neural Networks

out and receive activity from their associated nodes in other layers, but also

because each node is assumed to hold back some of its activity by an internal

memory mechanism.The model has been applied successfully to various psycholingusitic and

neurolinguistic data sets and can, for example, imitate patterns of speech er-rors made by normal speakers and aphasic patients. The network can imitate semantic errors, replacements of a word by another one close in meaning. In

the model, semantic errors are based on activity flow between lexical nodes

through their common associated semantic nodes. Phonological errors, re-

placement of words with similar sounds, can be based on phonological nodes

strongly associated with more than one lexical node. The model also explainswhy errors that are both phonologically and semantically related to the tar-

get word are far more frequent than one would predict on stochastic grounds

(Dell, 1986; Dell et al., 1997).

Different authors diverge in the exact architecture postulated for their

symbolic connectionist networks, with some proposing that more node types,

layers, and levels of processing are necessary. One influential proposal isthat apart from phonological, lexical, and semantic representations, there

should be a separate representation of a word’s syntactic properties, andthat this syntactic word node should be introduced between the lexical and

the semantic layers (Levelt, Roelofs, & Meyer, 1999).

A major point of divergence between proponents of the modular and

the interactive approaches to cognitive processing concerns the processingproperties of the systems of nodes and connections. As mentioned, the mod-

ular approach states that processing at each level is autonomous – that is, not

influenced by processes occurring at the same time at other levels – whereas

the alternative view is that there is continuous forward and backward activityflow. Furthermore, the strict modular view puts that activity flow is possible

only in one direction, therefore necessitating separate networks for model-ing speech production and perception. The discussion between modularist

and interactionist schools was quite vivid, with the outcome being that both

approaches can provide models for impressive data sets (Dell et al., 1997;

Norris, McQueen, & Cutler, 2000).

From a neuroscientific perspective, some assumptions immanent to mod-

ular models may appear difficult to anchor in brain matter. For example, theidea that activity should be allowed to flow in only one direction through

a network is in contrast with neuroanatomical observations that most of 

the corticocortical connections are reciprocal (Pandya & Yeterian, 1985;

Young et al., 1995). Also, the idea that the processing of different types

of linguistic information – phonological, lexical, and semantic, for exam-

ple – should take place without interaction would not have a necessary

Page 128: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 128/331

6.2 Symbolic Connectionist Models of Language 111

correlate in, for example, an associative memory model of word processing.

In such a model, strong links would be present, for example, between neu-

rons responding to acoustic features of a stimulus word and those sensitiveto visual features of objects to which the word refers. These links would allow

for fast information exchange between phonological and semantic parts of 

the distributed word representation. Having said this, it should be added thatthis by no means implies that a modular approach would be incompatible

with a neurobiological perspective.

The interactive perspective discussed here comes close to a cell assembly

model as proposed in Chapter 4. In both types of models, semantic and

phonological information about a word is linked through strong connections.In both model types, an activated representation maintains its enhancedactivity status for some time. The representation of a word is the union of 

neural elements spread out over different “layers,” or brain areas. There is

one important difference between the models, however. The connectionist

architecture includes nodes that represent and process only one lexical entry.

These word-specific neural elements are not necessary in the cell assembly

proposal. The coactivated neuron sets in primary sensory and motor cortices

are assumed to be held together by neurons in other areas, but neuronsinvolved only in the processing of one particular word would not be required.

Each neuron included in the functional web representing one particular word

could also be part of cell assemblies of other words as well. It is clear that,

in a computational architecture, word-specific nodes make the simulations

easier, because the activity level of the lexical node and its relation to thatof other lexical nodes provides a simple criterion for selecting one of the

words. In an associative network of overlapping functional webs, it may be

much more difficult to determine which of the assemblies is most active.Furthermore, activity excess is one of the dangers an associative memory

model runs into, a problem that is easily avoided in the symbolic network:

for example, a regulation mechanism within a lexical layer only allowingone node to be selected could be used. Therefore, the cell assembly model

requiresaregulationmechanismaswell(Braitenberg,1978a;Wickens,1993).

The regulation issue was addressed in Chapter 5.

Although the lexical layer may have computational advantages, one may

argue that such a layer of word-specific neural element is not an a priorirequirement within a neurobiological system. The functional web approachwould suggest that there are neuron sets in different primary cortical areas

that are activated, on a regular basis, when a given word is being processed,

and that there are neurons in other areas providing the connections between

these neurons. These additional neurons whose main function is to hold the

functional web representing a word together would respond to excitation

Page 129: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 129/331

112 Neural Algorithms and Neural Networks

of sensory and motor cells also being part of the ensemble. If each of thesebinding neurons was also be part of other cell ensembles, this may even

have advantages from a neurocomputational point of view. As mentioned inChapter 5, simulation studies of associative memories showed that the best

storage capacity is achieved if neuronal assemblies are allowed to overlap,

so that each neuron is part of several assemblies (Palm, 1982). Therefore,it is not necessary to assume word-specific neurons in a network in which

each word has its specific neuron set. However, these considerations do not

exclude the possibility that in a network of overlapping neuronal assemblies,

each representing one individual word, there might be neurons that are

part of only one assembly. Again, however, these cardinal cells (Barlow,1972) would not be required. Still, an important common feature between anapproach postulating functional webs and a symbolic connectionist network

is that each word is assumed to be represented as an anatomically distinct

and functionally discrete set of neurons or neuron-like elements.

6.3 Distributed Connectionist Models of Language

The distributed network type most commonly used to model language re-sembles symbolic networks because both network types are made up of lay-

ers of neurons and connections between layers. An important difference is

as follows: Most symbolic networks include so-called local representations –usually single artificial neurons – that can represent elementary features of 

the in- and output, but also more complex entities; for example, letters,

phonemes, and words. An example already discussed is the separate lexical

node postulated for each individual word. In contrast, distributed networks

use activity vectors specifying the activity of all neurons in one or morelayer(s) to represent words and other more complex linguistic structures. Inthis view, all neurons contribute to the processing of every single word. One

cannot say that only the active neurons of an activity vector would be rele-

vant for coding the respective word because changing the value of one active

neuron in the vector (to inactivity) and changing that of an inactive neuron

(to activity) may have a strong effect as well. This is particularly so if the

coding is not sparse, as is frequently the case in the so-called hidden layer of 

multilayer networks. Thus, each neuron contributes to the fully distributedrepresentation. One may therefore conclude that distributed connectionist

networks and localist symbolic networks reflect two extreme views at the

two ends of a continuum. Each individual neuron would be a “Jack of all

trades,” so to speak, in fully distributed approaches, but a master of almost

nothing (only one single word or phoneme, etc.) in the localist framework.

Page 130: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 130/331

6.3 Distributed Connectionist Models of Language 113

It is clear that both views could be correct descriptions of what actually is

the case in the brain, but it is also clear that there is room for alternative

proposals in between the extremes.The cell assembly framework offers such an alternative possibility be-

tween these extremes. Each neuron is envisaged to contribute significantlyto the processing of some words, but not necessarily to all of them. Each neu-

ron has a critical role in distinguishing between some lexical alternatives, but

may nevertheless be irrelevent for distinguishing between words to which it

corresponds. Not one cardinal cell and not the entire system would process a

single word, but a strongly connected neuron set, a functional web. A lexical

representation would be the union of neurons involved in processing themotor program, acoustic properties, and semantic features of a word.

A further difference between a cell assembly model on the one hand and

a fully distributed approach or a localist account on the other hand lies in

the structure of the assumed between-neuron connections. The distributed

connectionist networks most commonly used in language simulations, and

also their localist sisters, do not usually include direct excitatory connec-tions between the neurons in one layer. For example, an input layer in which

each neuron represents a phonological feature would leave these featureneurons without direct connections, although their activity may, in the ab-

sence of other active neurons, define a word. In contrast, a model of func-

tional webs based on neuroanatomical data must assume strong links within

cortical areas and between directly adjacent areas, and therefore betweenneurons related to sensory processes in a certain modality. This implies that

the different neurons sensitive to acoustic properties of a stimulus word

have a relatively high probability to be connected to each other directly,

rather than by way of intervening areas or layers. Because the active neu-rons defined by an activity vector are not connected to each other directly,

they do not form a functionally coherent system. This distinguishes themfrom cell assemblies or functional webs with strong internal links between

their neuron members, especially between their closely adjacent neurons.

However, given there are reciprocal connections between the layers of the

network, artificial neurons in different layers may strengthen their connec-

tions to neurons in other layers as a consequence of associative learning,

thereby indirectly linking the active neurons defined by individual activ-ity vectors. Still, the structural differences between the connectionist ar-

chitectures and a view incorporating more details about neuroanatomical

connnections remain.

The classical type of distributed connectionist network is called the per-

ceptron (Minsky & Papert, 1969; Rosenblatt, 1959). It consists of neurons

Page 131: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 131/331

114 Neural Algorithms and Neural Networks

arranged in two layers, the input and the output layers. Each neuron in

the input layer is connected to each neuron in the output layer. Activity

flows from input to output only. Connection strengths (or weights) vary asa function of associative learning between pairs of input–output patterns.

Rosenblatt (1959) proved that perceptrons can learn to solve only a certaintype of simple classification problem called linearly separable. Rumelhart,

Hinton, and Williams (1986) showed that a modification of the perceptron’s

architecture, through the addition of one additional layer of neurons, the

so-called hidden layer between the input and output layers, and by an ex-

tension of the learning rule, this limitation can be overcome. The networks

were now able to learn to solve more complex classification problems, forexample, classifications requiring the Either–Or operation, and this made

them much more interesting for cognitive scientists. It was mentioned in

Section 6.1 that a network of a certain type requires three neuronal steps to

solve the Either-Or problem (Kleene, 1956).

Three-layer perceptrons have been used successfully to model aspects

of language processing. For example, there are models that classify speechsignals (Waibel et al., 1995), others that mimic important aspects of the

infant’s learning of language specific information as described by elementaryrules (Hare, Elman, & Daugherty, 1995), and simulations of the effects of 

focal brain lesions on language functions (Hinton & Shallice, 1991; Plaut

& Shallice, 1993) and of the recovery of language functions after stroke. The

results of these simulations are not only of great theoretical relevance, butthey have found many useful practical applications as well.

Two- or three-layer distributed connectionist architectures do not have

memory. An input pattern activates the input layer and activity spreads to the

(hidden and) output layer(s). This network architecture must be modifiedfor addressing problems posed by syntax. To allow the network to process

long strings of words, it was necessary to introduce a memory mechanismin the architecture that allows the storage of information about past events

for a longer time span. Such a memory is necessary for assessing syntactic

relationships between temporally distant language units, as, for example, be-

tween the first and last words of a long sentence. Therefore, a memory layer 

allowing for reverberation of activity, thereby providing the basis of informa-

tion storage, was added to the three-layer architecture (Elman, 1990). Eitherthe hidden layer was given within-layer connections so that each of its neu-

rons fed back onto itself and the other neurons in the hidden layer, or loops

were introduced by connecting the hidden layer reciprocally to a separate

memory layer. Compared to three-layer perceptrons, these networks includ-

ing a memory layer could be shown to be more powerful in storing serial-

order relationships. They are capable of learning subsets of syntactically

Page 132: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 132/331

6.4 Hot Topics in Neural Network Research on Language 115

complex structures, for example, aspects of so-called center-embedded sen-

tences (Elman, 1990; Elman et al., 1996).

6.4 Hot Topics in Neural Network Research on Language

Among the many facets of connectionist research on language, work on two

topics can nicely illustrate in which way our understanding of the language

mechanisms have been furthered by simulation studies. The point here is not

so much that the actual simulation taught us important lessons. The critical

colleagues who consider many of the results of simulation studies as trivial

and emphasize the relevance of rigorous thinking about a problem, afterwhich a simulation might be close to redundant, may therefore be partially

correct. The point here is rather that simulation studies and comparable

theoretical work made scientists think about details of neuron circuits, and

this led to important insights.

Both issues to be addressed in this section are related to deficits in lan-guage processing, seen either in aphasia or in young infants before they are

fully capable of speaking their mother tongue. One issue is the explanation

of deficits in processing words of certain types. The other relates to younginfants’ acquisition of what one may want to describe as syntactic rules.

6.4.1 Word Category Deficits

Certain types of aphasia are surprising because the patients can produce

words of one type, but are unable to produce words of a second type. This

issue is addressed in Chapter 4 along with neuroimaging results. The princi-

ple observation that a word kind can be affected selectively by brain lesionhas first been made with regard to a disturbance called agrammatism (Pick,

1913). Agrammatic patients have difficulty producing words primarily char-

acterized by their grammatical function, such as articles, pronouns, auxiliary

verbs, prepositions, and also inflectional affixes. Geschwind (1974), a neurol-

ogist, recommended basing the diagnosis of agrammatism on a simple test:Ask the patient to repeat Sentence (8).

(8) No ifs, ands, or buts.

An agrammatic patient would usually not be able to repeat the sentence, andwould otherwise not be able to use the words included in it correctly. The

affected words are used mainly as grammatical function words and do not

refer to objects or actions. Difficulties with the respective word forms are

present even if these words are not actually used in their normal grammatical

function but instead, for example, in their nominalized forms (as nouns), as

Page 133: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 133/331

116 Neural Algorithms and Neural Networks

Sentence (8) illustrates. The abstract function words whose meanings are not

imageable at all are sometimes not the only word kinds affected. For some

patients, abstract nouns and verbs whose meaning is difficult to imagine arealso difficult to produce. A deficit in producing function words and other low-imageability words can be most pronounced in language production but can

be present in language perception as well (Pulvermuller, 1995; Pulvermuller,

Sedat et al., 1996). A deficit most pronounced for abstract low-imageability

words is also common in a neurological reading deficit called deep dyslexia

(Marcel & Patterson, 1978) and in semantic dementia (Patterson & Hodges,2001). One view on agrammatism and similar disturbances is that the im-

ageability of words is a crucial factor in determining their availability afterbrain lesion. Low-imageability words are more difficult than comparable

high-imageability words.

Even more surprisingly, however, agrammatism has a “mirror image” syn-

drome called anomia. Anomic aphasics do not have difficulty using functionwords and may even produce complex sentences without any problem but

cannot find some well-imageable content words from the lexical categories

of nouns, adjectives, and verbs (Benson, 1979).

How can the double dissociation between the deficit for low-imageabilitywords and function words in agrammatism and the deficit for the more con-

crete and imageable content words in anomia be explained? In the cell as-sembly framework discussed in Chapter 4, a highly imageable word would

be represented by a perisylvian cell assembly with strong connections to

neurons in other brain areas organizing referential aspects of the word, that

is, the motor programs and perceptual patterns to which it can refer. The cell

assembly would therefore be distributed over wider cortical areas, including

perisylvian areas but also other sensory- and action-related areas as well. Incontrast, a function word lacking any concrete associations that cannot beused to refer to concrete objects or actions would be represented by a func-

tional web without strong links to neurons in action- or perception-related

areas. Therefore, function words and also highly abstract and not imageable

content words would be realized in the brain by more focal functional webs

restricted to the perisylvian cortex. Figure 6.7 illustrates schematically the

postulated difference in circuits in the language-dominant hemisphere. A

further implication would be that the function words’ networks are morestrongly lateralized than are the networks representing highly imageable

content words. The word types are thus assumed to be realized as functional

webs with different cortical topographies (Pulvermuller, 1995).

Assuming more localized perisylvian functional webs for grammaticalfunction words and more widely distributed cell assemblies for concrete

nouns, verbs, and adjectives makes a few additional comments necessary.

Page 134: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 134/331

6.4 Hot Topics in Neural Network Research on Language 117

Figure 6.7. Schematic illustration of left-hemispheric distributions postulated for cell as-semblies representing high-imageability content words (left) and highly abstract functionwords (right).

First, it may appear more adequate to postulate such differential represen-

tation not at the level of individual words, but at that of meaningful units,

morphemes, instead. In this view, not entire content words, but “contentmorphemes” or the stems of content words, are represented (cf. Section 4.4).

(The reader may forgive the imprecision that the more common term, con-

tent word, is usually used in this book because it is much more common.)

Furthermore, there is no principle difference in the grammatical function

of a function word, such as the article “the”, and a functional affix, such asthe plural suffix “-s” attached to nouns. Their inclusion in a sentence usually

does not add semantic information to it, but may make the word string ei-

ther acceptable or unacceptable. The neuronal representations of functionwords and affixes could therefore exist without a semantic part. At least, they

can be conceptualized without a part including information about referen-

tial meaning that would be stored primarily outside the perisylvian region.In this respect, the functional webs representing different kinds of  function

items, words and affixes, may be similar.

A second point one may wish to raise is the following: The assumption

that no semantic information is being added to a sentence by the inclu-

sion of function items does not generally hold true for all members of this category. The regular past suffix “-ed” and the auxiliary verb form“was”, for example, include information about time, and it is clear that a

cortical representation of these lexical items must include this information

as well and bind it to the phonological information about the respective

form. However, it is less clear where in the brain this semantic information

characterizing some function words and affixes should be laid down. This

Page 135: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 135/331

118 Neural Algorithms and Neural Networks

information is not about these words’ referential semantics, and, therefore,

considerations such as the ones made in Chapter 4 are not relevant. By

postulating cell assemblies, or functional webs, for function words and af-fixes, it remains clear that the semantic information immanent in some of 

them must have a brain internal correlate, although it is doubted that thislocus of storage and processing is cortical and outside the perisylvian ar-

eas. In this context, it is relevant that the differential cortical processing of 

function words and stems of concrete content words receives support from

neurophysiological imaging studies (Brown, Hagoort, & ter Keurs, 1999;

Neville, Mills, & Lawson, 1992; Nobre & McCarthy, 1994; Pulvermuller,

Lutzenberger, & Birbaumer, 1995; Shtyrov & Pulvermuller, 2002a). In thecontext of studies reviewed in Chapter 4, these neurophysiological differ-

ences are best explained in terms of the difference in referential semantics

between content words and function items (for discussion, see Pulvermuller,

1999).

A parsimonious explanation of the double dissociation between agram-

matism and anomia becomes possible on the basis of the proposal that func-tion items are processed by strongly lateralized neuron ensembles restricted

to the perisylvian cortex, whereas content words have less lateralized cor-responding neuron ensembles distributed over various areas of both hemi-

spheres. Presumably, the likelihood of a word to be affected by a brain

lesion depends on the degree to which its word representation has been

damaged.A lesion restricted to the perisylvian regions, for example in Broca’s re-

gion in the inferior frontal lobe, would remove a high percentage of neurons

included in a function words’ webs. The concrete content words’ webs that

have their additional neurons outside perisylvian space (and also in the otherhemisphere) would not suffer so much, that is, a smaller percentage of their

neurons would be affected by the same lesion. This explains deficits primar-ily affecting function words and other abstract items, as seen in agrammatic

patients and related syndromes. The mirror-image pattern of deficits would

be expected if a lesion affected areas primarily outside the perisylvian re-

gion, for example, in the inferior temporal lobe, the temporooccipital region,

or various areas in the hemisphere not dominant for language. In this case,

only the concrete content words’ webs would be affected, thereby explaininga selective deficit in processing these items. This explanation of the double

dissociation between agrammatic and anomic deficits has been grounded

in neural network simulations and is in good agreement with data about

the cortical locus of lesions causing these disturbances (Pulvermuller, 1995;

Pulvermuller & Preissl, 1991). Focusing on deficits for low- vs. high-image-

ability content words in deep dyslexia, a similar account has been developed

Page 136: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 136/331

Page 137: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 137/331

120 Neural Algorithms and Neural Networks

relationships between irregular stems and past tense forms in an associative

manner (Pinker, 1997).

From the perspective of neural networks, however, one may ask whethertwo separate systems for rules and exceptions are actually necessary to han-

dle regular and irregular inflection. Rumelhart and McClelland (1986, 1987)showed that a two-layer perceptron can store and retrieve important aspects

of both past tense rules and exceptions. It can even produce errors typical

for children who learn past tense formation, such as so-called overgeneral-

izations (e.g., “goed” instead of “went”).

From a linguistic perspective, the two-layer model of past tense proposed

by Rumelhart and McClelland has been criticized because it did not appro-priately model the fact that rule-conforming behavior is by far most likely

to be generalized to novel forms. The past tense form of a newly introduced

verb such as “dif” will thus almost certainly receive an “ed” ending if one

intends to use it in the past tense (“diffed”). This is even so in languages

in which most verbs have irregular past tense forms and only a minority of 

the verbs conform to the rule. The rule is nevertheless used as the default,and generalized to novel forms and even to rare irregular items. This is a

problem for a subset of connectionist models because the strongest drivingforces in associative networks are the most common patterns in the input.

However, there are distributed three-layer networks that solved the prob-

lem of default generalization surprisingly well (Hare et al., 1995). An im-

portant determinant was that rule-conforming input patterns are maximallydissimilar whereas the members of an irregular class resemble each other.

Consider the different regular forms watch, talk, and jump in contrast to

the similar members of an irregular class sing, ring, and sting. Because the

regulars are so heterogeneous, they occupy a wide area in input space. Therepresentation in input space of a novel word is thus most likely to be closest

to those of one of the many different regular forms, and this is one reasonwhy so many new items are treated as regular by the network. However, if a

newly introduced item happens to strongly resemble many members of a reg-

ular class, for example, the pseudo-word pling, it is, in many cases, treated

as regular. These observations may lead one to redefine one’s concept of 

regularity: A rule is not necessarily the pattern most frequently applied to

existing forms, but is always the pattern applied to a most heterogeneous setof linguistic entities. The heterogeneity of the regular classes may explain

default generalization along with the great productivity of rules.

The simulation studies of the acquisition of past tense and other inflection

types by young infants suggest that neural networks consisting of one single

system of layers of artificial neurons provide a reasonable model of the

Page 138: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 138/331

6.4 Hot Topics in Neural Network Research on Language 121

underlying cognitive and brain processes. In this realm, the single system

perspective appears equally powerful as an approach favoring two systems:

one specializing in rule storage and the other in elementary associativepatterns.

Neuroscientific data and theories have shed new light on the issue of asingle-system vs. double-system account of rule-like behavior. The discovery

of patients with brain lesions who were differentially impaired in processing

regular and irregular past tense forms was important here. Patients suffering

from Parkinson’s disease or Broca’s aphasia were found to have more diffi-

culty processing regular forms, whereas patients with global deterioration of 

cortical functions as seen, for example, in Alzheimer’s disease or semanticdementia showed impaired processing of irregular forms (Marslen-Wilson& Tyler, 1997; Ullman et al., 1997). This double dissociation is difficult to

model using a single system of connected layers, but is easier to under-

stand if different neural systems are used to model regular and irregular

inflection.

Another argument in favor of a double system account comes from a neu-

robiological approach. Within the cell assembly framework, a typical verb

stem has its putative neurobiological correlate in a widely distributed func-tional web that includes neurons involved in the processing of actions and

perceptions related to the meaning of the verb. The to-be affixed or infixed

phonological material – for example, the ending “ed” or the [ae] inserted

into “ring” may initially be processed by a network restricted to perisylvianspace. The widely distributed assembly and the more focal perisylvian as-

sembly can be linked by two types of connections. One possibility is provided

by local connections within perisylvian areas, and the other by long-distance

connections between the verb stem’s assembly neurons far from the Sylvianfissure and the perisylvian neurons contributing to the processing of the past

tense form. The local and long-distance links have different likelihood, twoadjacent neurons having a higher probability of being connected than two

neurons in distant cortical areas (see Chapter 2). Assuming that most of 

the relevant connections between the two cell assemblies are indirect, that

is, through one additional neuronal step, the situation can be schematized

as in Figure 6.8. Two systems of connections, a high-probability pathway

linking perisylvian neurons and a low-probability pathway between neuronsoutside perisylvian space and neurons inside, would be available for storing

links between present and past tense forms. Given parameters are chosen

appropriately, the two low- and high-probability systems specialize differ-

entially in the storage of rules and irregular patterns. Similar to a two-layer

perceptron, the low-probability system is best at storing the simple mapping

Page 139: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 139/331

122 Neural Algorithms and Neural Networks

system 2

system 1

Figure 6.8. A neuroanatomically in-spired network architecture consistingof two systems of connections specializ-ing in exception learning and rule stor-age, respectively. The system at thetop includes a link with low connec-tion probability. The system at the bot-tom has high connection probabilitiesthroughout. Associative learning in thissystem leads to rule strorage primar-ily in the high-probability system (bot-tom) and to exception storage primar-ily in the low probability system (top).There is reason to assume that two typesof neuroanatomical links are relevant forthe learning of morphology. (For furtherexplanation, see text.) Adapted fromPulverm ¨ uller, F. (1998). On the matter of 

rules. Network: Computation in Neural Systems, 9 , R1–51.

between irregular present forms that resemble each other and their past

tense forms. In contrast, the complex classification problem, the mapping

between the heterogeneous regular stems and their past tense forms, is bestaccomplished by the three-layer component with high connection probabil-

ities. A full three-layer architecture is necessary for this, because Either-Or

relations are required for the classification of the heterogeneous items in

the regular class. When the two components are damaged differentially, the

network produces the double dissociation between regular and irregularinflection seen in neuropsychological patients. This approach explains theneuropsychological double dissociation while being consistent with aspects

of the acquisition of past tense formation by young infants (Pulvermuller,

1998). The explanation is based on principles of cortical connectivity and

has implications for the cortical loci of the wiring organizing rules and ex-

ceptions, respectively.

Page 140: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 140/331

6.4 Hot Topics in Neural Network Research on Language 123

Together, the neuropsychological double dissociation and the neurobio-

logical consideration argue in favor of a two-system model of regular and ir-

regular inflection. In contrast to a modular proposal stating that two systems

are concerned exclusively with regular and irregular processes, respectively,

the neuroscientific variant would suggest a gradual specialization caused by

differential connection probabilities. The fruitful debate between scientists

favoring single- or double-system accounts of rule-like knowledge reveals

the importance of multidisciplinary interaction between the linguistic, cog-

nitive, computational, and neurosciences.

Page 141: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 141/331

CHAPTER SEVEN

Basic Syntax 

Thischapterhighlightsafewproblemsofserialorderthatmayposeproblems

to a biologically realistic model, problems best explained in the context of 

theories designed to solve them. Therefore, the following paragraphs sketch

examples of linguistic descriptions of syntactic regularities. In addition, tech-nical terms are introduced. Occasionally, a sketch of what a neuron-based

approach to syntax might look like may intrude but is developed systemati-cally only in Chapters 9–12.

There are many different approaches to syntax in theoretical linguistics,

and it is not necessary to give an overview of them in the present context.

This chapter highlights examples, their choice being primarily motivated bythe historical development. So-called phrase structure grammars and their

offspring rooted in the work of Harris (1951, 1952) and Chomsky (1957) are

in the focus, because the superiority of phrase structure grammars to a model

of serial order in the McCulloch–Pitts (McCulloch & Pitts, 1943) traditionwas one of the main reasons in the 1940s and later to base syntax theories onthese more abstract algorithms rather than on neuronal algorithms. Apart

from approaches related to and building on phrase structure grammars, an

alternative framework whose roots also date back to the 1940s is mentioned

occasionally. This framework, or family of theories, goes back to Tesniere

and is called dependency grammar . Clearly, phrase structure grammars and

dependency grammars have been modified and much developed, but, as is

argued later in this chapter, some of their underlying main ideas persist.In addition to an introduction of syntactic terminology and theory, this

chapter reviews arguments in favor of and against different views on serial-

order mechanisms. Here, the main focus is on the question of what abstract

syntactic theories achieve more than neuron-based proposals. It is also asked

what the neuron-based proposals could offer what the abstract syntactictheories lack.

124

Page 142: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 142/331

7.1 Rewriting Rules 125

A linguist familiar with most syntactic theories may safely skip most of 

this chapter, but may want to glance over the list of questions that must be

addressed when building on a McCulloch–Pitts approach to syntax.

7.1 Rewriting Rules

Syntactic regularities are sometimes expressed in terms of  rewriting rules.A rewriting rule is a formula expressing that something can be replaced

by, or be rewritten as, something else. The rule consists of an arrow and

symbols to its left and right. The arrow can be read as “is rewritten as” or

“is replaced by.” A subtype of such rules, context-free rewriting rules, haveonly one single symbol to the left of the arrow and one or more symbols on

its right. Examples of context-free rewriting rules are given later. A set of context-freerewritingrulesdefinesa context-freegrammar .Asetofrewriting

rules can also be called a phrase structure grammar , or PSG.

Calling a set of rewriting rules a grammar implies a formal view on what

a grammar should actually achieve. This view restricts grammar to aspects

of the form of sentences, without consideration of the meaning of words

and strings and their role in communicative interaction. In more recent pro-posals, the interaction of formal syntactic rules and principles with semantic

processes is also considered. Although it appears justified to restrict consid-

eration to the question of how serial order of meaningful language units is

achieved, the task of a grammar may nevertheless be considered to cover

not only the form of sentences, but aspects of their meaning as well.

In modern linguistics, great breakthroughs have been achieved in the for-mal description of languages since the early 1940s. The early breakthroughs

were based on PSGs and more sophisticated grammar theories building on abasis of PSG rules (e.g., transformational grammars; Chomsky, 1963). Mean-

while, these theories have again been replaced in what is sometimes called

the standard view in linguistics (Chomsky, 2000; Haegeman, 1991). Still, how-ever, many modern syntax theories build on a kernel of PSG rules or a kernel

of principles and parameters that translate into language-specific context-

free grammars. When aiming at a neurobiological model of serial order, it

may be advisable to consider this common basis of many theories of syntax.

What exactly is a phrase structure grammar implementing a context-freegrammar or context-free rewriting system? As mentioned, it includes rewrit-ing rules that are subject to one restriction: Only one symbol is allowed to

the left of the arrow. At least one of these rules includes a starting symbol ,

abbreviated by an “S,” to the left of its arrow. S can be used to start the

generation process of a sentence. By applying a rewriting rule, the starting

symbol is being replaced by other symbols. These are symbols for syntactic

Page 143: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 143/331

126 Basic Syntax

categories, which are sentence parts, such as subject, predicate or object,

or noun phrase (NP) and verb phrase (VP). Additional rewriting rules can

specify the replacement of syntactic categories by other syntactic categories,or the replacement of syntactic categories by lexical categories, which are

word types sharing grammatical properties, such as noun (N), verb (V), ar-ticle/determiner (Det), or preposition (Prep). Symbols for lexical categories

finally can be replaced by symbols for actual words. Rules specifying the

insertion of words are called lexicon rules. Thus, after replacing symbols by

symbols and symbols by words, a word string or sentence results.

This can be illustrated by the following example rules:

(1) S NP VP

(2) NP (Det) N

(3) VP  V (NP)

(4) N {Betty, machine}

(5) V {laughs, cleans, switches}

(6) Det {the, a}

Starting with Rule (1), S yields NP and VP, a noun phrase and a verb

phrase. The noun phrase includes a noun, N, and can include an articleor determiner, Det, as well. This is the meaning of Rule (2), in which the

brackets to the right of the arrow indicate that the item within brackets, the

determiner, is optional. According to Rule (3), the VP can expand to a verb,

V, plus NP, but the NP is optional. The symbol for the lexical category N can

be replaced by a real noun. Candidates are listed in Rule (4), which states

that one of the words within curved brackets can replace the N symbol. Rules(4)–(6) are lexicon rules because they replace the abstract lexical category

labels by lexical elements. In contrast, Rules (1)–(3) are syntactic rules in thenarrow sense; they replace syntactic category symbols by other symbols for

syntactic or lexical categories and are not concerned with individual words.

By inserting (4) into (2) and (2) into (1), the string (7) can be derived.

(7) Betty VP

By additional insertions of (4) and (6) into (2) (yielding “the machine”), and

of (2) and (5) into (3) (resulting in the VP “cleans the machine”), sentence

(7) can be rewritten as (8).

(8) Betty cleans the machine.

Other correct grammatical sentences can also be generated based on the

small set of rules. They include the sentences (9) and (10), but also strange

word strings such as (11).

Page 144: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 144/331

7.1 Rewriting Rules 127

(9) Betty laughs.

(10) The machine cleans Betty.

(11) The Betty laughs machine.More generally, a phrase structure grammar, as one possible description

of a context-free grammar, includes

(i) symbols representing lexical categories (e.g., N, V),

(ii) symbols for syntactic categories (e.g., NP, VP),(iii) rewriting rules with one syntactic or lexical category symbol to the

left of the arrow (e.g., NP Ar N), and

(iv) a starting symbol (S).Phrase structure grammars specify sets of lexical and syntactic categories,

or word and phrase types, that differ between languages. Modern syntactic

descriptions attempted to replace such idiosyncrasies of individual languages

by a more universal formulation of lexical categories, phrase types, and rules

(Haegeman, 1991; Jackendoff, 1977). Phrase types are now defined with ref-erence to their main lexical category, their head. Instead of verb and verb

phrases, different projection levels of the verb are being distinguished. This

makes the formulation of grammars of individual languages more econom-ical. The equivalents of phrase structure rules can be derived from more

abstract principles by setting a few language-specific parameters. As an ex-ample, the abstract principle might be that each word from a major lexical

category X requires a complement (and forms a so-called X projection to-

gether with it), allows for an adjunct, and also requires a “specifier” to form

a “maximal projection” (X). The language-specific parameters to be set

include, for example, the information about whether the specifier has to be

placed before or after the verb. The syntactic categories would thus be con-sidered derivatives, or projections, of one of the major lexical categories.

Instead of VP, one would write V, or Vmax. This indicates that separate

syntactic categories, independent of the major lexical categories they are

driven by, are not necessary in a formal grammar [cf. (ii)]. The more sys-

tematic treatment of rules (on the basis of principles and parameters) and

syntactic categories (considered to be projections of lexical categories) donot affect the general arguments made here.

At this point, it may be difficult to see why a network of neurons storingand processing syntactic information, called a neuronal grammar , should not

be capable of solving the same tasks as a PSG does. One may speculate that

a neuronal grammar could include

(i) neuronal elements representing morphemes and words

(ii) directed connections between neuronal representations

Page 145: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 145/331

128 Basic Syntax

(iii) criteria for activity dynamics in the network and for string accep-

tance

and that these neuronal elements and dynamics could be as powerful indetermining serial order of neuronal excitation as are formal grammars in

determining the serial order of words and meaningful language elements.

However, it has been argued that this is incorrect, that a neuronal devicecannot, for principled reasons, process string types that can be handled easily

by a formal model of syntax.

7.2 Center EmbeddingThe beauty of a context-free grammar lies in the fact that rules are recur-

sively applicable. By repeatedly using rules, one can obtain syntactically

correct sentences of considerable complexity. As pointed out by Chomsky

(1963) and Levelt (1974), a neuron network of the McCulloch–Pitts type is

equivalent to a regular grammar  or right-sided linear grammar . A context-

free grammar is more powerful than a regular grammar; it includes recursiverules that allow one to produce sentences in which other sentences are em-

bedded, in which even further sentences can be contained. This “Russian

doll” game can be played, as has been argued by linguists, with, in principle,

no upper limit for the number of recursive operations, so that sentences of 

mild to extreme complexity can be synthesized. The fairly complex sentences

would be characterized by center embedding, if the same sentence types areembedded in each other, by self-embedding.

Examples of center-embedded sentences or strings are listed as (12)–(14).

Brackets have been inserted to help to structure the strings.

(12) Betty gives the machine (that had been cleaned the other day) to Bob.

(13) The rat (the cat (the dog chased) killed) ate the malt.

(14) Anyone (1 who feels (2 that (3 if so many more students (4 whom we

haven’t actually admitted 4) are sitting in on the course (4 than ones we

have 4) (4 that the room had to be changed 4) 3) then probably auditors

will have to be excluded 2) 1) is likely to agree that the curriculum needs

revision.

Clearly, the two more extreme word strings, (13) and (14) (the latter being

taken from Chomsky, 1963), are confusing and improbable. Sentence (13)

is not easy to understand, and paper and pencil are necessary for work-

ing out a possible meaning for (14). Nevertheless, the argument maintained

by some theoretical linguists would be that the internal grammar mecha-

nism itself would, in principle, be capable of generating and processing such

Page 146: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 146/331

7.2 Center Embedding 129

strings, were there not the theoretically not interesting limitations to mem-

ory and processing capacity. In fact, the “English sentence” (14) was used by

Chomsky “to illustrate more fully the complexities that must in principle beaccounted for by a real grammar of a natural language” (Chomsky, p. 286,

1963).Example strings (12)–(14) show that there is, in fact, an upper limit for the

language processing device in the human brain. The fact that much effort is

needed, and even pencil and paper, to make sense out of (14) demonstrates

that the brain-internal grammar mechanism is limited with regard to this

analysis. There appears to be no empirical basis for a brain-internal grammar

mechanism capable of dealing with strings such as Sentence (14) withoutexternal aids.

There are languages other than English in which center embeddings are

more common. The German translation of (13) might be considered much

less irritating by many native speakers of German than its translation, sen-

tence (13), is by native speakers of English.

(15) Die Ratte [die die Katze (die der Hund jagte) get ¨ otet hat] ass den Malz

auf.

Thus, there can be no doubt that center embeddings are possible, at least in

certain languages. For a realistic model of language, the relevant question is

what the upper limit for the number of embeddings is. Empirical research

on this issue indicated that three embedded sentences are at the upper limitand are already difficult to understand (Bach, Brown, & Marslen-Wilson,

1986). This study also revealed that speakers of Dutch understand sentences

including multiple crossed dependencies better than do speakers of Germanwith center-embedded sentences. The authors argued that this rules out a

pushdown mechanism (cf. p. 131f) as a universal basis of the human parsing

mechanism. However, it may well be that a pushdown mechanism is avail-able, but a distinct mechanism supporting crossed dependencies is available

as well, and is more effective under conditions of high information load

(Pulvermuller, 1994).

An abstract description of the problem posed by multiple center embed-

dings is as follows: In many languages there are strings such as (16) thatlater require complement strings in the reverse order, as schematized by(17). Here, the upper-case letters represent words, morphemes, or larger

constituents including words. An initial string part would therefore require

a complement string whose constituents exhibit the mirror-image sequence

of the initial sentence part. If (16) is the initial part, (17) would be the mirror

image part, and (18) would be the center-embedded construction.

Page 147: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 147/331

130 Basic Syntax

(16) AB . . . M

(17) M . . . BA

(18) AB . . . MM . . . BA

Here, all pairs I, I are assumed to be complements of each other. If an

element I has a complement I, it means that if I occurs in a string, I must

also be present. In (18), the sequence of constituents is therefore followed

by the mirror-image sequence of their complements. Sentence (13) can bedescribed by the mirror-image order of syntactic categories in (19).

(19) NP1 NP2 NP3 VP3 VP2 VP1

Here, each sentence consists of two complements, one NP and one VP. Num-bers indicate the sentence to which the constituents belong.

To understand a center-embedded sentence such as (12), (13), or (15), the

parts of each sentence that belong together must be related to each other.

To complete sentence (13) after the third subject noun phrase has been

uttered, not only information about the missing constituents of the mainand subordinate clauses must be kept in memory, but information about

their serial order as well.

Context-free grammars offer to deal with the phenomenon of multi-

ple center embedding in the following way. A mirror-image string such as

(20) can be generated by starting the generation process with the starting

symbol, and by rewriting it as indicated in (21), then applying Rule (22), andfinally Rule (23). (The optional “s” would be left out only in the latter rule

application.)

(20) A B C C B A

(21) s A (s) A

(22) s B (s) B

(23) s C (s) C

(24) B A C C A B

(25) A A A A A A A A

In these and similiar examples, lower-case letters, a, b, c, . . . , are used for

syntactic category labels and upper-case letters, A, B, C, . . . , are symbols for

actual parts of a string. Application of Rules (21)–(23) in a different ordermay result in (24), and repeated application of (21) in (25). Please note thatthe strings and Rules (20)–(25) are more abstract and simpler than the earlier

examples taken from, or applicable to, real languages.

PSG Rules (21)–(23) would express that syntactic category symbols

(indicated here by small letters) can be rewritten as strings of lexical category

symbols (capital letters) and other, in this case optional (brackets), symbols

Page 148: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 148/331

7.2 Center Embedding 131

for syntactic categories. After the arrow, the syntactic category symbols can

appear to the left or right of a lexical category label. This is an important

feature of context-free rewriting systems. If a syntactic category symbol ap-pears only on one side of a syntactic category symbol, and if rules with more

than one syntactic category label to the right of the arrow are not permitted,the descriptive power of the grammar is reduced. Such a one-sided linear

grammar then, for example, is not able to process center-embedded strings

(see Levelt, 1974). In right-sided linear grammar, the rules have their lexical

category symbol always to the right of the lexical category symbol, and are

therefore of the form (26).

(26) a B c

Again, the small and large letters refer to syntactic categories and string

parts, respectively. As mentioned, right-sided linear grammars are also calledregular grammars and are equivalent to networks of McCulloch–Pitts neu-

rons. Thus, it appears that the McCulloch–Pitts devices lack something

context-free grammars exhibit.

The memory mechanism postulated to be necessary for the processing of a context-free grammar is called pushdown memory. Context-free grammars

have been shown to be equivalent to pushdown automata, one type of linear-

bounded automaton. In contrast to a finite automaton of the McCulloch–

Pitts type, which is characterized by a finite number of neuron elements,

each with a finite number of possible activity states, a pushdown automatonincludes a pushdown store of unrestricted capacity. It is called pushdown

memory because it is accessed in a first-in last-out manner, like a stack.

New information is entered on top of the stack, above the older entries,and memory access is from top to bottom. Whatever piece of information is

being entered into the store first is read out only after all information put into

the store later has been removed. Conversely, the last piece of informationentered into the store is read out first.

Having a pushdown store available makes it easy to store the knowl-

edge about the mirror-image sequence of the verb phrase complements

in example sentences (13) or (15), or the complements in the strings (19),

(20), (24), and (25). Whenever an initial element is being perceived or pro-duced, information about its complement, as specified by the respective rule[e.g., (21)–(23)] can be entered into the memory stack. The complement

entered when the last constituent of the initial sentence part is being used

remains on the top of the stack. Only after it has been read out is it possible

to access the information about the complement of a constituent appearing

earlier in the initial part of the string. This continues until the information

Page 149: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 149/331

132 Basic Syntax

about the complement of the initial constituent of the sentence has been read

out. The way the pushdown automaton stores and accesses its information

thus produces the mirror-image sequence of complements.Both a regular and a context-free grammar, a McCulloch–Pitts device and

a pushdown automaton, can generate, or specifically respond to, strings of arbitrary length. It is sometimes claimed that syntax algorithms are special,

because they allow the generation or acception of, in principle, an infinite

number of possible strings based on a finite set of rules. However, this is also

true for regular grammars and their equivalent, McCulloch–Pitts neuron

ensembles. This is explained in the context of neurons specifically responding

to events of arbitrary length, the All and the Existence neuron in Section 6.1.Another feature sometimes considered to be specific to syntax algorithmsis the recursive applicability of their rules. However, this is, again, also a fea-

ture of regular grammars. An example of a simple abstract regular grammar,

which is analogous to the context-free grammar (21)–(23), follows. Clearly,

the rewriting Rules (27)–(29) also define a grammar, in this case a right-

sided linear grammar, in which each rule can be applied recursively in theprocessing of a sting.

(27) s A (s)

(28) s B (s)

(29) s C (s)

These rewriting rules can generate, for example, strings of A’s, B’s, or C’s of 

any length, or a sequence in which these three symbols are mixed arbitrarily.The grammar is called right linear because, after the arrow, all syntactic cate-

gory symbols (s) are to the right of lexical category symbols (capital letters).

Analogs in terms of McCulloch–Pitts circuits are obvious. For example, aneuron activating itself can produce an unlimited sequence of activations of 

the same type. (if certain abstractions are being made).

One of the string types regular grammars cannot process are center-embedded strings of indefinite length. It has been argued by theoretical

linguists that center-embedded strings are characteristic of many languages

and that it is, therefore, important for grammar theories to allow for their

processing without upper limit for the number of embeddings allowed. Note

that this argument depends crucially on the assumption that extremely com-plex center-embedded strings such as (14) are relevant in grammatical pro-cessing. From this perspective, it appears that context-free grammars are

superior to regular grammars with regard to the description of natural lan-

guages. Regular languages, and their equivalent neuronal automata, were

therefore considered inadequate as descriptions of the syntax of natural lan-

guages. The neuron-based devices were abandoned by linguists, if they had

Page 150: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 150/331

7.2 Center Embedding 133

at all been considered in the first place. A door between the language and

brain sciences was closed, or possibly never opened.

This historical move is surprising because not too long after the elab-oration of context-free PSGs and syntactic theories building on them, it

could be shown that neuron-based models can be made equivalent to PSGsin describing context-free rewriting systems. The major difference between

a McCulloch–Pitts network, with its finite number of possible states, and

the pushdown automaton, with its stack of indefinite depth, obviously lies

in their storage capacity, which is limited in one case but unlimited in the

other. It is trivial that a device with unlimited storage capacity can outper-

form a device with limited capacity on certain tasks. Therefore, the obviousmove was to think about possibilities to upgrade finite automata. For ex-

ample, recursive transition networks were proposed and shown to be equiv-

alent to PSGs. These formed the basis of a new class of grammar devices,

called augmented transition networks (see also Section 6.1; Kaplan, 1972;

Winograd, 1983; Woods, 1973). A different approach was rooted in the idea

that McCulloch–Pitts networks could be allowed to grow so that the numberof their neurons, and therefore the amount of information they can store,

is unlimited (Petri, 1970; Schnelle, 1996b). It is unclear whether this view isrealistic as a description of brain-internal processes. However, it is uncon-

troversial that a McCulloch–Pitts network with unlimited storage capacity

and therefore a putatively infinite number of possible states is equivalent

to a Turing machine, and is therefore even more powerful than a pushdowndevice (Schnelle, 1996b). It appears that the view about the inferiority of 

McCulloch–Pitts approaches to syntax cannot be maintained.

In summary, the limited vs. unlimited storage capacities of the respective

devices is not an important difference between approaches to serial-orderprocessing in the McCulloch–Pitts tradition on the one hand and those build-

ing on abstract pushdown automata on the other. Not only can this differencebe compensated for easily by extending neuronal networks, its relevance is

also questionable, given the dubious status of unacceptably complex strings

with multiple embeddings such as (14). As mentioned, empirical evidence

(Bach et al., 1986) suggests that speculations about unlimited pushdown

stores are not required.

Perhaps a more important difference lies in the way pushdown automataand McCulloch–Pitts networks treat center-embedded sentences of finite

length. Clearly, a McCulloch–Pitts automaton can also store information

about a finite number of constituents that occur in the input. If a network is

confronted with a string ABC, neurons each specifically activated by one of 

the string elements could store that the respective elements were present.

The mechanism could be the same as those underlying the Existence neurons

Page 151: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 151/331

134 Basic Syntax

described in Section 6.1. However, although in this case information about

the elements would be stored, the information about their serial order would

be lost. One way to store the information about the sequence would be byway of sequence detectors – as also discussed in Section 6.1 – that respond

only if input elements have been active in a particular order. However, for thethree elements A, B, and C, there would, theoretically, already be 33 – that is,

27 possible sequences – each requiring its specific mirror-image complement.

Postulating so many different sequence detectors may render the proposal

unrealistic and therefore unattractive. It appears important to think about

how a realistic neuronal device could store information about the important

properties of such sequences of elements without the need for introducingnew sequence detectors for each and every possible series of constituents.

One approach is suggested by the pushdown storage mechanism: Perhaps

each of the neuronal elements is active and some property of their activation

retains the information about their serial order. This idea may appear attrac-

tive, because it allows for more economical information storage. However,

it certainly needs elaboration, in particular with regard to possible realisticneuroscientific under pinnings. This is discussed in Chapter 12.

7.3 Discontinuous Constituents and Distributed Words

Is it possible that a neuron-based approach to syntax is more elegant for

describing aspects of serial order in language than classical linguistic ap-

proaches rooted in the tradition of PSGs?

In Section 7.1, an example grammar was presented including six rewriting

rules, (1)–(6) (see p. 126). Whereas these rules allow for the generation

of several strings, for example, (9)–(11), they do not allow for generatingsentence (30).

(30) Betty switches the machine on.

The problem here is that the last word’s placement would not be specified

by any of the earlier rules. At least one rule would need to be modified. For

example, Rule (3), which is repeated here as (31) for convenience, could bereplaced by (32):

(31) VP  V (NP)

(32) VP  V (NP) (on)

Given Rule (32) is being used, sentence (30) can be generated. The verb

could then be followed not only by the noun phrase (“the machine”) but

by the verb particle (“on”) as well. The downside of this formulation in

(32) is that the intimate relationship between the two words switch and on

Page 152: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 152/331

7.3 Discontinuous Constituents and Distributed Words 135

disappears in this representation. The relationship between these two words

is so close that the two can actually be considered one single composite

word. This would in no way be a strong assumption. There are languages, forexample, German, in which (33) is translated as a single word (34), which

can also appear in its distributed form (35).

(33) switch . . . . . . on

(34) einschalten

(35) schalten . . . . . . ein

Theproposaltocallitemssuchas(33)–(35) distributed words datesbackto

Harris (1945), a linguist. Distributed words or morphemes are a special caseof  discontinuous constituent , grammatically closely related sentence parts

that are separated from each other in time in spoken language and sepa-

rated in space in written language. Discontinuous constituents allow for, or

can even require, that other material be inserted between the parts of the

distributed representation.The example of distributed words is even more interesting than that of 

discontinuous constituents at the level of larger syntactic categories. The

main parts of a main clause, noun and verb phrases, separated from eachother by a relative clause in center-embedded sentences, such as (12)–(15),

would be an example of discontinuous constituents above the level of words.

As discussed in Section 7.2, center embeddings can easily be dealt with bycontext-free rewriting systems. This is not so for distributed words such as

(33) and (35). The reason is that the word is usually assumed to be dealt with

as a whole. The grammar would process the word as a whole although the

different parts occur at different places in the sentences and are separated in

time or space. In a PSG, a lexicon rule would specify the two word parts and,in the generation process of a sentence, this lexicon rule would be applied

only after the VP has been broken down into V and NP [Rule (3)]. Thus, one

would ultimately end up with a string of words in which verb and verb particle

are next to each other, although they may actually appear, for example, on

opposite sides of an NP in the sentence.

A sentence including a discontinuous constituent or distributed word re-mains grammatical if the string inserted between its parts is made longer and

longer. This is illustrated by examples (36)–(39), in which the two parts of the discontinuous constituent appear in italics. (The examples have already

been mentioned in Chapter 6 and are repeated here for convenience.)

(36) Betty switched  it on.

(37) Betty switched  the coffee machine on.

(38) Betty switched  the nice brown coffee machine on.

Page 153: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 153/331

Page 154: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 154/331

7.3 Discontinuous Constituents and Distributed Words 137

To generate strings (42) and (43), the following rewriting rules (which

form a one-sided linear grammar) may be used:

(46) a A b(47) b switches c

(48) c B d

(49) d C (on)

Starting with a, the sequential application of Rules (46)–(49) yield string

(42) or (43), depending on which of the two alternatives allowed by (49),

with or without particle, is being chosen.

The parentheses in (49) suggest that the choice of the final particle is anoptional process. However, this is not appropriate. Selection of the particle

must be coupled to the selection of one of the two readings of the verb.Notably, the two verbs switch and switch...on in (42) and (43) have different

meanings: (43), but not (42), implies that there are two machines, one of 

which is being replaced. There is therefore motivation to distinguish two

different verbs that share the phonological form switch.

A tentative idea would be to represent the two parts of a distributed word

in a neuronal grammar by using bistable elements, for example, the Exis-tence cells discussed in Chapter 6. The two parts of a distributed word could

be implemented by two Existence cells that are in some way functionally

related. If one of the two is activated, the other one would also be aroused

to some degree, and thereby prepared to accept the second part of the dis-

tributed word. The bistable nature of the Existence cells would guarantee

that time lags of some length can intervene between the two complementword parts. This tentative idea would also need to be worked out in detail.

The problem of distributed words and other long-distance dependencies isvirulent not only for one-sided linear grammars, but also for any context-free

rewriting system. In order to solve it, different solutions have been proposed.

For example, the two complementary constituents have been assumed to be

generated together and one of them transported away from the other toits correct position by a movement or transformation (Harris, 1945, 1951,

1952, 1957, 1965). In sentence (44), the verb particle on would therefore be

moved away from the verb root to its sentence final position. The idea that

the grammatically correct sentences of a natural language are the result of movements of constituents within strings first“base-generated” by a context-

freerewritingsystemhasbeenworkedoutingreatdetailandinvariousformsby Chomsky (1957, 1965).

A different solution of the problem of discontinuous constituents requires

introduction of new categories of nonterminal nodes. One example is the slash features in modified PSGs (Gazdar, Klein, Pullum, & Sag, 1985; Joshi

Page 155: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 155/331

138 Basic Syntax

& Levy, 1982) that specify syntactic objects lacking a complement or are

otherwise incomplete. According to this proposal, the information about a

missing verb particle could be, so to speak, transported through the sentencerepresentation. The nodes of the syntactic tree intervening between the twocomplements (e.g., verb root and particle) would incorporate the informa-

tion that a particular type of complement (e.g., the particle) is missing.

Still a different solution of the problem posed by discontinuous con-

stituents and long-distance dependencies is offered by a more fine-grained

categorization of words into lexical subcategories, as first proposed in the

context of valence theory and dependency grammar. Assume we wish to

represent the root of the particle verb and the verb without particle as twodistinct elements in the lexicon, say switch—1 (root of transitive verb re-

quiring the particle, abbreviated here as “V14par”) and switch—2 (“plain”

transitive verb, “V14”). The differentmeanings of the word switchesmaypro-

vide sufficient motivation for introducing two word representations in this

case. (The idea of more fine-grained lexical subcategorization also underliesthe dependency rules formulated in the next section.) This would have the

consequence that the rewriting system would need to become more complex

than in (46)–(49). These rules and the syntactic categories they operate onmust be doubled to distinguish between the representations triggered by the

two different verbs.

(50) b switches—1 c1

(51) c1 B d1

(52) d1 C on

(53) b

switches—2 c2(54) c2 B d2

(55) d2 C

Technically speaking, it is now guaranteed that, if switches—1 is chosen

by application of Rule (50), the particle on must follow at the end of the

sentence. However, choice of the other variant of the verb by Rule (53)

Rules out a final verb particle on. This makes the syntactic description more

precise at the cost of introducing more material, lexical categories, and rules,into the right-linear (regular) grammar, in this case.

However, there is still nothing in the one-sided linear grammar that would

allow for representing the fact that it is precisely the verb stem and the

particle that are interdependent. This disadvantage could be overcome, for

example, by using one of the strategies outlined previously (making use

of movement or slash features) or by a neuronal machinery implementing

Page 156: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 156/331

7.4 Defining Word Categories in Terms of Complements 139

the close relationship between complements by direct links between their

representations and bridging the time delay by bistable neuronal elements

(i.e., Existence neurons, memory cells).Other disadvantages of the more complex right-sided linear grammar

introduced for solving the problem of discontinuous constituents and dis-tributed words are that so many different rules are now necessary and also

that the number of symbols for syntactic categories was doubled. If there

were several discontinuous constituents or long-distance dependencies in a

sentence, the number of additional rules and nonterminals would multiply. If 

the number of long-distance dependencies is n, the number of nonterminals

and the number of rewriting rules would increase with 2

n

. This exponentialincrease is not attractive. In a neuronal circuit in which discontinuous con-

stituents are realized as linked bistable units, the number n of elements to

be kept in memory can be represented by n memory circuits.

This example may indicate that, although finite neuronal automata of the

McCulloch–Pitts type and one-sided linear grammars are equivalent, there

may be syntactic phenomena aspects of which can be coded more easily, ormore elegantly, in a neuron network than in a rewriting system.

7.4 Defining Word Categories in Terms of Complements:Dependency Syntax 

In PSGs, rules were proposed to operate on syntactic categories. A differ-

ent approach is to base grammar algorithms on lexical categories (or wordclasses) and their complements, the necessary additions they require, and

adjuncts, the optional additions they allow for. This alternative approach

was taken in the 1950s by a syntactic theory, or class of theories, calleddependency grammar (Tesniere, 1953, 1959).

Dependency grammars can be formulated in terms of dependency rules

that operate on lexical categories and do not necessitate syntactic categories.These rules specify that the presence of an element of lexical category A in

the string requires the presence of additional elements from lexical cate-

gories B, C, D, . . . . The latter elements are therefore said to be dependent

on A.

To distinguish rewriting and dependency rules from each other in thepresent context, a different notation for the latter is used. Hays (1964) andGaifman (1965) proposed to abbreviate the valence and complement struc-

ture of verbs by brackets in which symbols for the required complements

are inserted either to the left or right of an asterisk symbolizing the posi-

tion of the element outside brackets they are dependent from (for a more

elaborate formalism, see Heringer, 1996). Here, a similar notation is used

Page 157: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 157/331

140 Basic Syntax

with an asterisk between slashes representing the position of the parent cat-

egory and the labels of the dependent categories between brackets. Optional

rules, that is, rules that specify possible but not required adjuncts, are shownin double brackets. Curly brackets frame alternative words that can replace

labels for lexical categories. The context-free grammar (1)–(6) has its analogin the dependency grammar (56)–(65):

(56) V14par (N1 /∗/ N4 Par)

(57) V14 (N1 /∗/ N4)

(58) V1 (N1 /∗/)

(59) N1 ((Det /∗/))

(60) N4 ((Det /∗/))(61) V14par {switches}

(62) V14 {cleans}

(63) V1 {laughs}

(64) N {Betty, machine}

(65) Par {on}

This example of dependency grammar includes five syntactic dependency

rules, (56)–(60), and five lexicon rules, (61)–(65). This formulation is slightlymore detailed than the formulation chosen in (1)–(6) because it includessubcategories of verbs and nouns. In the present grammar fragment, nouns

are either subjects, and are therefore in the nominative case, N1, or they are

accusative objects, N4, and are therefore in the accusative case. The verbs

are intransitive “one-place” verbs, V1, which do not require complements

except for their subject noun, or they are subclassified as V14, transitive

verbs requiring an additional accusative object, or, as a third possibility, as

transitive particle verbs V14par, which require a verb particle in additionto an N1 and N4. The lexicon rules also become more detailed, consistent

with the larger number of subcategories. Rule (64) applies for both N1 and

N4. The increase in sophistication of the grammar and the larger number of 

rules now allow for a more precise adjustment of the grammar to the set of 

strings the grammar should generate.Here are illustrations of the meaning of the dependency rules: (56) would

read “a transitive particle verb stem requires a noun in the nominative case

that precedes it and a noun in the accusative case and a particle that follow it.”Rules (59) and (60) state that nouns require determiners that precede them.

The dependency system (56)–(65) allows for the generation of some of 

the example sentences discussed previously and are repeated here for con-venience. Following each sentence, its structure is illustrated on the basis

of the syntactic dependency rules. The dependency rule appropriate for the

respective verb type is first selected and other rules are inserted to specify

additional relevant lexical categories.

Page 158: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 158/331

Page 159: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 159/331

142 Basic Syntax

Figure 7.1. The structure of a sentence is illustrated based on different approaches togrammar. (Upper left) Context-free phrase structure grammar is used to analyze the string.(Upper right) X-bar syntax. (Lower left) Dependency grammar. (Lower right) A symbolismthought to represent neuronal elements and connections between them. (For explanation,see text.)

(70) Betty switches the machine on.

As mentioned, analysis of this string is not a simple task. Obviously, one

problem is that it includes a particle verb that is realized as two separate

constituents that belong together, switches and on. Note that in the trees,abbreviations of lexical and syntactic categories have been changed slightly

to make them even shorter, and Vp replaces Vpar and Ar is used instead of 

Det. Also, the verb suffix is analyzed (Vs).

The diagram on the upper left in Figure 7.1 gives an example of a syn-

tactic tree constructed from a simple context-free grammar or PSG. Themacroscopic sentence structure is analyzed according to Rules (1) and (3)(Section 7.1), and the noun phase is broken down as detailed in Rule (2). In

addition, the verb is rewritten as verb followed by a verb suffix. The analysis

leaves open how the verb particle Vp should be linked into the construction.

As mentioned in Section 7.3, one possibility would be an indirect link to the

verb phrase so that a line connects the VP label to the Vp(ar). As mentioned,

Page 160: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 160/331

7.5 Syntactic Trees 143

this would obscure the relationship between the verb root and its particle

and would therefore not be optimal. The verb and its particle can be linked

to each other directly only through a line if projection lines cross. This is,however, not allowed in this kind of graph, which is, as we know, conceptual-

ized as strictly two dimensional and therefore without crossings of projectionlines. A third possibility is to attach the verb suffix to the noun or noun phrase

nodes directly and define the intervening constituents, the VP and NP (and

N), as incomplete constituents missing a verb particle. In this case, the feature

that the particle is missing could originate from the verb stem representa-

tion and be transported through the intervening nodes of the tree, finally

reaching the constituent directly connected to the particle representation.ThetreegraphontheupperrightofFigure7.1iscongruenttothePSGtreeon the upper right. The only difference is the relabeled nodes. Each syntactic

category label has been replaced by the name of a lexical category plus prime

signs indicating the distance between the corresponding nodes. This is similar

to an analysis based on X-bar theory. The charming thing here is that it

becomes apparent that the syntactic categories are closely related to lexical

categories. The problem with the attachment of the verb particle remains.

The diagram on the lower left of Figure 7.1 shows how the same stringmay be represented by a dependency grammar lacking nonlexical syntactic

categories entirely. The tree is near equivalent with the bracketed repre-

sentation in (69). The lines in this tree have a different meaning than in

the two earlier graphs. Apart from relationships between lexical entries andlexical categories, they express the dependency relationship between lexi-

cal categories. In this formulation, it is possible to attach the verb particle

to the verb it belongs to without violating the projectivity principle. This

is not to say, however, that the projectivity problem is absent in this gram-mar type for other discontinuous constituents or unbounded dependencies

(see Heringer, 1996). A major difference distinguishing the dependency ap-proach from rewriting systems is that in the former the syntactic properties

of one of the words in the sentence, in this case the verb, determines the

structure of the sentence representation.

An ensemble of connected neuronal elements that may serve similar pur-

poses as the aforementioned trees is sketched on the lower right of Figure 7.1.

It is based on the idea that in the brain, there are specific representations of lexical categories. In this case, symbols for neuronal connections replace the

lines indicating rewriting or dependency relations in the tree graphs.

There appear to be no a priori reasons for preferring one of these rep-

resentations, although it appears at first glance that treatment of the verb

particle is least problematic in the representations at the bottom. However,

again, it is well known that after introduction of additional principles into

Page 161: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 161/331

144 Basic Syntax

PSG-based rewriting systems, aspects of the problem of discontinuous con-

stituents and unbounded dependencies can be solved (Gazdar et al., 1985).

Furthermore, note that none of the representations in Figure 7.1 cap-tures the fact that there is agreement between the subject noun Betty and

the verb suffix -es – the phonological realization of the present third personsingular feature of the verb. For the graphs to describe such fine-grained

structural relationships, some refinement would be necessary. A possibility

is again offered by transporting features through the tree representation so

that the feature can be shared by lexical elements distant from each other

in the string. In the case of the agreement between subject noun and verb

suffix, the feature “third person present singular” could climb through thetree, from the noun in subject position to the verb. Clearly, this would re-quire a mechanism different from the one expressed in terms of rewriting or

dependency rules or related descriptions of serial order in sentences. One

may ask whether such an additional feature transport mechanism is required

(Pulvermuller, 2002).

7.6 Questions for a Neuronal Grammar

Of course, there are many open questions that must be addressed by a neu-robiological approach to serial order, such as

(a) How can center-embedded strings be represented (e.g., those of theform ABCCBA)?

(b) How can discontinuous constituents and distributed words be real-

ized (e.g., switch . . . on)?

(c) How is it possible to specify the syntactic relatedness of distantelements in a string (e.g., noun and verb agreement)?

These problems have been dealt with in great detail in syntactic theories,

but a brain perspective on them has not been offered, or at least not been

specified in great depth.More questions pose problems to a neuronal device designed to process

serial order, one of which is the possibility to store the fact that the same lex-

ical entry or syntactic category ocurs more than once in a sentence. Clearly,one word can occur repeatedly in a sentence, sometimes even with the same

meaning.

(71) The dog chased the cat.

(72) Die Ratte [die die Katze (die die Hunde jagten) get ¨ otet hat] ass den

Malz auf. (Engl.: The rat [whom the cat (whom the dogs chased) killed]

ate the malt.)

Page 162: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 162/331

7.6 Questions for a Neuronal Grammar 145

In sentence (71), the determiner the occurs twice, whereas in example (72),

the German word die is used five times – twice as a relative pronoun (i.e.,

that 

andwho

), and three times as definite articles marked for casus, nu-merus, and genus. It is clear that repeated use of words in single sentencesis a rather common phenomenon, and that words are frequently used in

the same grammatical function and semantic meaning when being repeated.

The problem that a representation is being processed twice also occurs if a

syntactic category is used repeatedly in a derivation.

Context-free rewriting systems offer to deal with the phenomenon, as

discussed in detail in the context of examples (20)–(29). Rules are usually

defined as being recursive, and there is no restriction on the number of occur-rences of syntactic categories, lexical categories, or lexical items in one tree.

In contrast, however, there are models for which an account of this feature is

difficult. Localist connectionist models of language, for example, implement

words as nodes in a network. If two occurrences of the same word in one sen-

tence are to be modeled, it may be necessary to activate the representationtwice simultaneously, so to speak, a possibility not offered by most models.

Nevertheless, repeated access to the same representations – words, lexical or

syntactic category representations, and/or rules – and storage of this multi-ple use is necessary for modelling important properties of language (Sougn ´ e,

1998).This type-token problem is not easy to realize within a neuronal machine.

Section 6.1 discusses one form of repeated processing of the same infor-

mation when modeling a device responding to a string of several identical

stimuli or events. However, this solution has the disadvantage of not keeping

track of how many times the neural representation has been activated. It is

clear that to produce or understand sentences (71) and (72) successfully, onemust know not only that a string of several the’s or die’s was present, but

also exactly how many of the’s or die’s were present to answer the question

of whether this string was grammatical and acceptable; the same knowledge

is necessary for comprehending the string successfully.

To realize repeated use of the same word or lexical category, one may

postulate several neuronal representations of such an entity. For example,predicates have been proposed to be available a finite number of times in

neuronal automata (see Shastri & Ajjanagadde, 1993). Processing of sev-eral tokens of the same type would then correspond to activation of two,

three, or more neuronal devices. One may assume to have, for instance, five

McCulloch–Pitts neurons representing the word the. Although this assump-

tion is certainly possible, it is not very attractive because (i) it is difficultto explain why the brain should invent several representations of the same

entity, and (ii) the choice of any upper limit for the number of representa-

tions would be arbitrary.

Page 163: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 163/331

146 Basic Syntax

A second possible view posits that different levels of activity of one neu-

ronal device could code for multiple processing of the same linguistic ele-

ment. This solution, however, requires that activity levels coding 1, 2, andmore tokens of the same type are distinct. Given the gradual loss of activity

over time seen in many memory cells (see Chapter 2), this does not appearto be a fruitful strategy. Information about one or several occurrences of the

same element may easily be mixed up if neuronal representations exhibit

exponential activity decrease over time. There is cross-talk and, therefore,

loss of information.

As a third perspective on solving the problem of multiple use of one

linguistic entity, one may assume that multiple activity patterns in the sameneuronal device realize presence of several tokens of the same type in a string

or in the linguistic representation of a string. This possibility is explored

further in Chapter 12.

Another question posing problems for a serial order network is related to

the organization of lexical categories. All syntactic theories use lexical cate-

gories, and there is little doubt that some kind of a category representationmust be postulated in a neuronal architecture of serial order as well. How-

ever, how this could be done is not clear. In some of the proposed neuralnetworks, individual neurons or neuron ensembles are assumed to repre-

sent and process individual words or morphemes (see Chapters 4–6). For a

network underlying the production and perception of syntactic strings, it is,

for obvious reasons, desirable not to have separate representations for eachchain of words; otherwise each word string would have to be learned sepa-

rately, and the ability to generate new grammatical sentences could not be

explained. Therefore, it appears advantageous to have a representation of 

what different words and morphemes have in common. If there is a neuronalrepresentation of these common features of the words and morphemes of 

a particular type, rules of serial order could be realized as the connectionsnot between individual word/morpheme representations, but as the con-

nections between the representations of common features of several words

(Pulvermuller, 2000).

To sum up, in addition to the problems (a) to (c), questions (d) and (e)

must be addressed:

(d) How can repeated use of the same word or lexical category within a

sentence be modeled and stored?

(e) How can lexical categories be realized in a neuronal network?

Tentative answers to questions (a)–(e) are proposed in Chapters 9–12.

Page 164: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 164/331

Page 165: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 165/331

148 Synfire Chains as the Basis of Serial Order in the Brain

connection from α to β. A single neuron α could, therefore, by way of its

direct projection to neuron β, arouse it whenever active.

However, it is unlikely that single cortical neurons connected in this wayplay a role in language processing. As mentioned in Chapter 2, the connec-

tions of most neurons in the cortex are known to be weak so that input fromone single neuron would usually not be sufficient to strongly enhance the

firing probability of a second neuron on which the first projects (Abeles,

1991). Data from both in vivo and in vitro intracellular recordings show

that individual postsynaptic potentials are small and unreliable (e.g., about

0.2 µV+/− 0.1 µV in the neocortex of the rat). Thus, the concept of strong

connections between individual neurons is not realistic. Near-synchronousinput through many of its afferent synapses is usually necessary to arouse a

neuron. It appears therefore to be more likely that sets of neurons project

onto each other, thereby making up broad neuron chains that determine

spatiotemporal patterns of activity. Waves of synchronous excitation could

then spread along these broad neuron chains.

Studies on the correlation of the firing in multiple neuronal units pro-vided strong evidence for complex spatiotemporal activity patterns in which

many neurons participate (Abeles et al., 1993; Vaadia et al., 1995). The firingprobability of a single neuron could best be determined when more than one

preceding neuronal event and behavioral context were taken into account.

This context dependence cannot be modeled based on a chain of single neu-

rons, each projecting onto the next in the chain. If this view were correct,one would expect each neuron to have a constant influence on the firing

probability of other neurons. Instead, it was found that two neurons firing

with a particular exact delay were good predictors of the firing of a third

neuron. In contrast, each of the former neurons firing alone or both neuronsfiring with a slightly different delay would not have a detectable effect of the

third one. Thus, the firing of a neuron was found to be strongly dependenton the context of more than one other neuronal events.

The complex context dependence of firing probabilities follows from a

model in which groups of neurons are connected in chains. In this case, the

synchronous activity of one of the groups that are connected in sequence

is necessary to arouse the next set. This type of neuronal circuit has been

labeled a synfire chain (Abeles, 1991). Figure 8.1 illustrates a simple modelof a synfire chain. In this illustration, each neuron needs two simultaneous

inputs to become active, and each of the sequentially connected sets of the

chain includes three neurons. This is a simplification made for ease of ex-

hibition; the number of neurons of each neuron set connected in sequence

is probably higher, between 50 and 100 neurons (Diesmann, Gewaltig, &Aertsen, 1999), and their firing threshold is probably in the order of 5–10

Page 166: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 166/331

8.1 Neurophysiological Evidence and Neuronal Models 149

Figure 8.1. A synfire chain consisting of five subgroups of neurons connected in sequence.Circles represent neurons and v-shapes connections between cells. Each neuron is assumedto have a threshold of 2. Each subgroup activates only the next subgroup if all of its neuronsfire simultaneously. If the synfire chain is active, the neurons α, β, and γ  embedded in thechain will fire in a fixed temporal order (e.g., β 10 ms after α, and γ  another 30 ms after β).If α and β fire with this fixed delay, 10 ms, the probability for γ  to fire another 30 ms lateris enhanced. If α or β fire alone, and therefore independently of the synfire chain, then thefiring probability of γ  is not affected.

simultaneous inputs (Abeles, 1991). In the model circuit, simultaneous acti-vation of the three neurons on the left leads to a wave of activity spreading

to the right. The serially connected subsets, or steps, of the synfire chain ac-

tivate each other after exact delays. The neurons of each subset activate the

neurons of the next subset at exactly the same point in time.

Figure 8.1 explains the relationship between the assumed synfire mech-anism and the mentioned data on context dependence of neuronal firing.

Note that neurons α and β in this example circuit can be active indepen-

dently of each other. Given the signals they send out are weak, each of thembeing active alone would have only a negligible influence on neuron γ  at the

end of the chain. However, if they are active in a given temporal order, for

instance β 10 ms after α, this would likely be because the chain as a whole isactive, in which case γ  would be activated at a fixed delay after β, perhaps

30 ms later. Thus, the firing probabilities depend strongly on the context of 

other neuronal firings. The conditional probability of  γ  firing 40 ms afterα and 30 ms after β is high, but the firing probabilities of γ  in other contexts

for example, after α has fired but not β, may be low.The synfire model implies that a cortical neuron can be part of several

synfire chains and that it can therefore be a member of different spatiotem-

poral firing patterns. The individual neurons would then become active, on

a regular basis, in different well-defined behavioral and neuronal contexts.

In multi-unit studies of cortical activity, it was found that much of a neuron’s

activity that would otherwise be considered to be “noise” or “spontaneous

Page 167: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 167/331

150 Synfire Chains as the Basis of Serial Order in the Brain

activity” could actually be described in terms of frequently occurring spa-

tiotemporal firing patterns (Abeles et al., 1993, 1995).

To further illustrate the synfire mechanism, a schematic representation of two intersecting synfire chains is shown in Figure 8.2. Whenever the neuron

group at the upper left is active, an excitation wave spreads downward,terminating at the lower right. Although the neurons in the very center are

also heavily connected to the neuron groups at the lower left, activity on

the lower left dies out in this case. In the same way, a wave from the upper

right spreads only to the lower left. There are two distinct spatiotemporal

patterns of activity that are prevented from getting mixed up by the very

nature of the connections, although the structural bases for these patternsoverlap. Two of four of the neurons in the central layer, where the two synfire

chains cross and the two ovals overlap, belong to both synfre chains. Each

of the other neurons included in the synfire chains may also be part of other

chains as well.

An essential feature of the synfire model is that information highways

share part of their structural basis, which has the obvious consequence thateach neuron’s firing depends on the firing context of several other neurons.

As mentioned, the shared part of the two chains sketched in Figure 8.2 arethe neurons in the middle of the central neuron group, where the ovals

intersect. These neurons would become active as part of an activity wave

Figure 8.2. Synfire chains that cross. Eachcircle represents a neuron and y-shapesrepresent connections between neurons.Each neuron is assumed to have a thresh-old of 2. Possible phonemic correlates of subsets of the synfire chains are indicated.

Page 168: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 168/331

8.2 A Putative Basis of Phonological Processing 151

starting at the upper left, but would also be activated if an activity wave

started at the upper right. In this case, the firing of these neurons shared

by the two synfire chains alone would not decide the direction in which theexcitation wave progresses. However, given these shared neurons are active,

the direction of activity flow can be determined if more of the context of the firing of these neurons is taken into account. The left- and right-most

neurons in the central group, which are included in only one of the ovals,

have the role of  context indicators channeling the wave of activity either

to the left or right. Synfire chains can thus have shared neurons, in which

case context indicators are crucial for keeping them and the spatiotemporal

activity pattern they determine separate.

8.2 A Putative Basis of Phonological Processing

Synfire chains have been proposed as a putative neuronal basis of articu-

latory programs (Braitenberg & Pulvermuller, 1992). The exact timing of 

nerve cell firings determined by the circuitry would allow for realizing pre-

cisely timed articulations. From a cognitive perspective, the beauty of the

synfire chain mechanism lies in its potential to provide a straightforwardsolution to what Lashley (1951) described as one of the main aspects of 

the problem of serial order in behavior . If each phoneme or letter was re-

presented as a separate entity, the possible words of a language could not

be modeled simply by direct connections of 50 or so phoneme or letter

representations. Too many different sequences would be possible whenevera set of phoneme or letter representations are selected. If a set of repre-

sentations is activated, for example, the phonemes [t], [æ] and [b], there

would be no information about their serial order, so that different sequenceswould be permitted (e.g., “tab” and “bat”). However, if not phonemes but

context-sensitive phoneme variants were represented instead, each possi-

ble sequence could be realized by direct links between neuron sets. Thecontext-sensitive variant of phonemes would be, for example, [b] at word

onset followed by [æ] – which can be abbreviated #Bæ – [æ] following [b],

and followed by [t], bÆt, and [t] terminating a word and preceded by [æ],

æT#. The three context-sensitive phoneme-like units #Bæ, bÆt, and æT#

would determine the elements of the phoneme sequence and their serial or-der. A similar solution to the serial-order problem has earlier been suggestedby Wickelgren (1969).

IncontrasttoWickelgren’sapproach,whichacknowledgesonlyphoneme-

context effects among adjacent phonemes, it is well known that the phys-

ical properties of the realization of language sounds can depend on more

distant context. The synfire model would suggest that for each frequently

Page 169: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 169/331

152 Synfire Chains as the Basis of Serial Order in the Brain

encountered word, a synfire chain is being established and that the synfire

chains of different word forms intersect, cross, and overlap depending on

how thephysical features of the sounds

are being shared between the usualrealizations of the word forms.

In phonology and phonetics, context-dependent variants of phonemes aresometimes called allophones. Allophones are phonotactically determined

variants of a language sound with complementary distributions, the vari-

ation being determined by their position relative to other phonemes in a

string. There are also free variants of phonemes without obvious phonotactic

determinants and similar distributions (e.g., the [r] and [R] sounds generated

by the tip vs. back of the tongue as they coexist in German and French). It ispossible that the context-sensitive variants of a phoneme that are laid down

in the synfire chain include phonetic features of both phonotactic phoneme

variants and of the free phoneme variants the individual happened to be

frequently exposed to. The synfire model allows for specifying a putative

underlying mechanism of the processing of language sound variants.

It may be worthwhile to illustrate the processing of phonological andphonetic information in a synfire architecture. Figure 8.2 can be used for

this purpose, although the relevant articulatory–phonological mechanismsare much simplified. The focus is on articulatory mechanisms only, although

analogous mechanisms must be postulated on the acoustic side.

If the synfire chain starting at the upper left and running to the lower

right is considered the neuronal correlate of the syllable [bæt], its compo-nent neuron groups can be taken as the correlate of the relevant linguistic

elements, the language sounds as they are realized in the respective con-

text. Each sound representation would be composed of two different kinds

of neuronal elements: One related to invariant properties of the articula-tion of a phoneme and realized by the shared neurons of different chains;

the other related to variations of the articulation depending on the con-text sounds in which a given language sound is being articulated, realized

by the context-indicator neurons mentioned previously. These would cover

features of allophones, the context variants of phonemes.

This can again be made more plastic on the basis of Figure 8.2. The neu-

rons shared between the two representations of context-sensitive variants of 

the phoneme [æ] – the two middle neurons in the central layer of Figure 8.2shared by both chains – could relate to articulatory distinctive features of 

that phoneme (e.g., lips open but not rounded, tongue at the bottom of 

the mouth). In contrast, the neurons deciding between the possible suc-

cessors and distinguishing between the alternative synfire chains – the left-

and rightmost neurons of the central neuron group, the context indicators –

would process information about the articulatory movements determined

Page 170: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 170/331

8.2 A Putative Basis of Phonological Processing 153

by the context, the preceding and following articulations. Information about

differences between allophones is stored and processed here.

In this view, the neurobiological equivalent of a phoneme in contextwould therefore consist of neurons related to articulatory distinctive features

(the shared neurons) and others realizing context-sensitive phoneme variants

allophones (the context indicators).

Any discussion of a possible neurobiological correlate of phonological

processing must include some limitations of the proposal. Clearly, the synfire

model postulates exact delays. Although this is an advantage for modeling

precisely timed articulations, the model circuits may provide not enough

flexibility for explaining certain obvious phenomena. For example, speechrate can be high or low. If synfire chains were designed to model different

articulation speeds (i.e., rapid or slow pronunciation of a syllable), it would

be necessary to allow them to be tuned so that the basic time scale on which

they operate can be changed. One possibility is to assume that the chains

operate at different levels of background activity affecting all neurons in the

chain to a similar degree. Weak background activity slows the speed withwhich activity runs down the chain. The candidate explanatory process is that

more time is needed for temporal summation of neuronal activity necessaryfor activating each step in the chain. Higher background levels of activity

may speed up the chain because less temporal summation would be needed.

It is clear that this proposal must be worked out in more detail, but it is

equally clear that such elaboration is possible (Wennekers & Palm, 1996).In assuming that synfire chains realize the articulation of phoneme strings,

one may suggest that each of the functional webs underlying word process-

ing (see Chapter 4) could include a synfire chain controlling the articulation

of the word form. The functional web representing the phonological wordform would include, or would have attached to itself, a chain of neurons

whose spatiotemporal sequence of activity is timed precisely. Ignition of thefunctional web could therefore include or be followed by the sequence of 

neuron firings determined by the articulatory synfire chain. Ignition of the

word-related functional web would not necessarily also activate the syn-

fire chain. Such activation of the articulatory program may require that the

respective neurons be in a preparatory state.

This proposal does not address the fact that words can be realized indifferent ways, some of which are substantially reduced in rapid speech.

For example, the function word will  (but not its noun homophone) can be

reduced to ll . To model this in a network, one may attach two or more

mutually exclusive synfire chains to the representation of the function word.

As an alternative, one may model the reduced articulation on the basis of background activity levels. If articulation speed is very high, the articulatory

Page 171: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 171/331

154 Synfire Chains as the Basis of Serial Order in the Brain

network may be under strong background activation, so that the respective

chains become too fast for the neuromuscular articulatory machinery to

follow their commands. This results in omissions of sounds, in particular incontexts in which fast complex articulatory movements would be required

(e.g., “it comes” realized without the [t]).Keeping in mind the limitations of the proposal, it still appears that the

synfire model may allow for a fruitful approach to phonological brain pro-

cesses. Because it offers (i) a simple solution for one aspect of the serial

order problem, (ii) a mechanism for precisely timed articulations, and (iii)

a mechanism for coarticulation effects, the synfire model may provide a

neurobiological perspective on articulatory mechanisms. Clearly, this doesnot prove the model correct. There are also fruitful approaches to memory

processes for phonological sequences that exploit alternative, or possibly

complementary, mechanisms (Page & Norris, 1998).

One of the features this proposal shares with recent psycholinguistic ap-

proaches (Marslen-Wilson & Warren, 1994) is that it does not require sep-

arate representations of phonemes. Overlapping sets of neurons related tophonetic distinctive features and context features are proposed to corre-

spond to the contextual variants of each language sound. The intersection,or, as an alternative, the union of these overlapping neuron sets, can be

considered the (putative) neurobiological basis of a phoneme. No addi-

tional “phoneme nodes” are required; rather, several overlapping context-

sensitive representations of a phoneme coexist. Furthermore, feature over-lap of all context vatiants of one phoneme are not required. The family

resemblance argument (Wittgenstein, 1953) may hold for phoneme vari-

ants as it holds in the semantic domain (Section 5.2.3), and the underlying

neuronal mechanisms may be similar. Although the intersection of context-sensitive representations underspecifies the articulatory movements neces-sary for realizing the phoneme, each of the context-sensitive representations

itself overspecifies the articulatory movements of the phoneme because it

also includes redundant information about phoneme variants whose fea-

tures may, strictly speaking, not all  be necessary for distinguishing word

meanings.

8.3 Can Synfire Chains Realize Grammar?

It is tempting to apply the synfire model to higher-order sequences of mean-

ingful units, morphemes, and words. One may want to define and neuronally

implement a word’s syntactic role on the basis of its context words, the items

that frequently occur before and after it in continuous speech, and pos-

tulate a representation of these various contexts by multiply crossing and

Page 172: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 172/331

Page 173: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 173/331

156 Synfire Chains as the Basis of Serial Order in the Brain

of a noun or personal pronoun predicts the later occurrence of a

complement verb, but material may intervene between the two, as,

for example, in “Peter comes to town,” “Peter, the singer, comes . . . ,”“Peter, the greatest singer in the world, comes . . . ” (see also examples

in Chapter 7). A synfire model would not allow for such extremelyvariable delays.

These points are closely related to the issues discussed in Chapter 7.

Although they are generally difficult to account for in a neurobiologically

realistic model, the synfire model appears to be unable to solve them. One

may speculate whether it might be possible, in principle, to modify the syn-fire mechanism so that it could operate at a larger time scale and deal with

the various problems listed. However, the concept of synfire chains would bestretched in this case, and its most important feature – the temporal precision

they provide – would need to be removed [cf. point (5)]. There is no reason

for such redefinition of the concept. Rather, it is advisable to characterize

the mechanisms necessary for grammar processing and how they differ from

the synfire mechanism.

8.4 Functional Webs Composed of Synfire Chains

In Chapter 2, functional webs were defined as strongly connected neuronensembles. This does not have further implications about in which way the

assemblies are structured internally. One could envisage these networks to

be unstructured lumps of neurons in which average connection probability

and strength are high, but in which activity flow is not determined otherwise.

In Section 8.2, it was proposed that functional webs may include synfirechains for certain tasks (e.g., for defining articulatory programs). Takingthis view, one may still consider cell assemblies to be largely unstructured

neuron ensembles in which only subgroups of neurons exhibit well-timed

spatiotemporal activity patterns.

An alternative view is the one proposed by Donald Hebb (1949) in his

book on cell assemblies. He suggested that the cell assembly is actually a

reverberatory chain of neurons, with many reentrant loops through which

activity waves can travel repeatedly.This view about the inner structure of cell assemblies is illustrated in

Figure 8.3, which is taken from Hebb (1949). In this diagram, arrows are

used to represent neuron groups and the numbers are used to indicate the

sequence in which the different neuron groups are activated, given the cell

assembly is in a certain state of activity. If the cell assembly is active, this

Page 174: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 174/331

Page 175: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 175/331

158 Synfire Chains as the Basis of Serial Order in the Brain

(Pulvermuller, Lutzenberger et al., 1999). One may therefore speculate that

these high-frequency dynamics reveal the signature of functional webs, in-

cluding many reverberatory loops.

8.5 Summary

Synfire chains consisting of sets of neurons connected in sequence are a

well-established neurobiological mechanism of serial order. Each synfire

chain defines a precise spatiotemporal pattern of neuronal activity. Synfirechains may overlap and cross with other chains while keeping separate the

spatiotemporal activity patterns they organize. It is possible that the brainexploits the synfire mechanism for controlling precisely timed movements,

including articulations. Serial order of phonemes may be organized by syn-

fire chains built into or attached to functional webs representing words.

A mechanism different from that of synfire chains may be necessary forestablishing serial order of meaningful language units in sentences.

In summary, a neurobiological approach to serial order in language

processing suggests that different mechanisms underlie the processing of 

phoneme sequences within syllables and words on the one hand and the pro-cessing of word and morpheme sequences in sentences on the other hand.

Chapter 9 discusses the latter.

Page 176: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 176/331

Page 177: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 177/331

160 Sequence Detectors

cells, α and β, looking at adjacent areas A and B of visual space, a moving

stimulus first appearing at A and later appearing at B stimulates the neuronsα

andβ

sequentially. A third neuron,γ 

, receiving input from bothα

andβ

,may function as a detector of a movement in the AB direction. It should

respond to the sequential stimulation of  α and β, but not to the reversesequence. The mechanism yielding sequence sensitivity may involve low-

pass filtering of the signal from α, thereby delaying and stretching it over

time. Integration of the delayed and stretched signal from α and the

actual signal from β yields values that are large when the activation of  α

precedes that of β, but small values when the activations of α and β occur

simultaneously or in reverse order. This mechanism of directional sensitivitywas first described in the visual system of insects (Reichardt & Varju, 1959;

Varju & Reichardt, 1967). Analogous mechanisms of movement detection by

mediated sequence processing were uncovered in the visual cortex of higher

mammals (Barlow & Levick, 1965; Hubel, 1995), and a related mechanism

of sequence detection exists in the cerebellum (Braitenberg, Heck, & Sultan,

1997).The cerebellar mechanism appears to be closest to the mechanism of 

sequence detection suggested by McCulloch and Pitts’s model circuits(cf. Chapter 6). Actual delay lines, realized as the parallel fibers of the

cerebellum, ensure that precisely ordered sequences of neuronal events can

simultaneously stimulate the sequence-detecting cardinal cells, realized as

cerebellar Purkinje cells (Braitenberg et al., 1997). This is very close to themechanism sketched in Figure 6.2. The three input units in the figure would,

in reality, correspond to inputs to different parallel fibers. The delay units

would not be present in the cerebellar mechanism, but parallel fibers of 

different length would replace them. The sequence detector cell would beanalogous to a Purkinije cell responding only if several action potentials ar-

rive simultaneously in its flat dendritic tree through several parallel fibers.It would appear that many aspects of the mechanisms envisaged by McCul-

loch and Pitts could be confirmed by modern neurophysiological research

(see Heck, 1993, 1995).

Again, the specific feature of the mechanisms of mediated sequence pro-

cessing is that a sequence of elementary events is detected by a third-party

element (labeled γ  here) that receives input from the neuronal correlatesof the elementary events (labeled α and β). This mechanism of  mediated

  serial-order processing is in contrast with the unmediated or direct serial-

order mechanisms as, for example, synfire chains – because it is characterized

by the existence of neuronal elements devoted to computing serial-order in-

formation and mediating between more elementary units. Please note thatthe two possibilities – mediated, or indirect, sequence detection based on

Page 178: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 178/331

9.2 Sequence Detectors for Word Strings 161

extra third-party neuronal elements vs. direct sequence detection made

possible by the synfire architecture – had already been sketched and con-

trasted in Section 6.1 (cf. Figs. 6.2 and 6.3, respectively). Meanwhile, it be-came clear that there is strong evidence for the existence of both mechanisms

in the nervous system.

9.2 Sequence Detectors for Word Strings

A basic proposal explored in Chapters (10–12) of this book is that mediated

sequence detection may be relevant for processing the serial order of words

and morphemes in sentences. However, word-order detection cannot beachieved by exactly one of the mechanisms found in the visual system of 

arthropodes and vertebrates, or in the cerebellum, because of the time do-

main differences [cf. caveat (1) in Section 8.3]. As is the case for the synfirechain mechanism, the mechanisms for direction-sensitive movement detec-

tion usually apply for delays much smaller than 1 second, whereas sub-

stantially longer delays occur between sequentially aligned words and mor-

phemes. Still, however, Barlow and colleagues reported that some neurons

in the visual system of vertebrates exhibit rather long decay times (Barlow,Hill, & Levick, 1964), which could be compatible with the detection of se-

quences spanning tens of seconds. In addition, a model building on sequence

detectors fed by word webs can also be subject to all of the points raised

previously against a synfire model of word sequencing. Points (2) to (5) in

Section 8.3 and also Point (1), are therefore be addressed again. The strat-

egy is to explore what the mediated sequence detection mechanism already

well established by neuroscientific research can achieve, and how it would

operate at the level of functional webs.

(2) Numberof represented sequences : One may argue that a model based

on indirect sequence detectors for word strings requires a very largenumber of such detectors, each responding to a particular sentence.

However, this is not necessarily so. Like movement detectors, word-

sensitive sequence detectors can be assumed to operate on pairs of 

elementary units. If there is a sequence detector for each frequently

occurring sequence of two words, the number of necessary sequencedetectors can be reduced substantially. Still, the number would belarge [but see (3)].

(3) Categorization: If a sequence detectorγ  responds to a sequence “firstα1, then β1” of neuronal events, it is possible that it responds to a

sequence “firstα2, then β2” as well (whereα1, α2, β1,andβ2 symbolize

word webs). By connections toα1,α2, . . . , αm ontheonesideandto β1,

Page 179: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 179/331

162 Sequence Detectors

β2, . . . , βn on the other, γ  can be sensitive to activation sequences

of elements of groups of word webs – that is, to a sequence of any

member of theα

group followed by any member of theβ

group.Theα group could, for example, be the lexical category of nouns, and

the β group could be the verbs. The sequence detectors could operateon webs representing words and morphemes from given lexical

categories.

(4) Generalization: Suppose a sequence detector γ  is frequently active

together with the activation sequence of word webs α1 and β1, and

develops, by associative learning, strong connections to both so that

it finally responds reliably to the sequence “first α1, then β1.” Addi-tional confrontation with the sequences “first α1, then β2” may also

strengthen the sequence detector’s connections to β2, and finally, if 

theactivationof α2 isfrequentlyfollowedbythatof β1,theα2 webmay

furthermore be chained to γ . The “generalization” that the sequence

detector is also sensitive to the event “first α2, then β2,” although this

particular sequence may never have been present in the input, followsfrom this learning history. This type of substitution-based associative

learning can account for at least one type of generalization of syntac-tic rules to novel word strings. It requires that a few standard strings

be perceived frequently, and that substitutes replace members of the

string selectively so that their representation can be chained to thesequence detector.

(5) Variable delays: A sequence detector does not require fixed tem-

poral delays between the activation of the units feeding into it to be-

come active. Motion detectors of the Reichardt–Varju type

(Reichardt & Varju, 1959; Varju & Reichardt, 1967) can respond tostimuli moving with variable speed, and in the same way, a functionalweb fed into by two word webs may respond to their serial activation

independently of the exact delay in between their respective activa-

tion. A noun–verb sequence detector may therefore become active

whenever confronted with one of the strings “Peter comes to town,”

“Peter, the singer, comes . . . ,” or “Peter, the greatest disk jockey of 

the world, comes . . . .” Clearly, there must be an upper limit for the

delays possible, which, in a Reichardt–Varju inspired model, woulddepend on the decay times of the word webs and the characteristics

of the low-pass filter.

Returning, finally, to Point (1), the time-scale problem, it appears that de-

lays of several seconds do not constitute a principled problem for a model of 

mediated sequence detection. Again, the decay time and the characteristics

Page 180: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 180/331

9.3 Sequence Detectors and Syntactic Structure 163

of the assumed low-pass filter may determine the time scale on which se-

quence detection is possible. If functional webs corresponding to words feed

into a sequence detector, the activation and decay times of these word webswould likely be crucial. At present, there are no data revealing these dy-

namics of the neurobiological correlates of words. However, one may allowfor educated guesses. Relevant neurophysiological data come from memory

cells (see Chapter 2) whose dynamics may reflect the activity status of dis-

tributed memory networks they are part of. If this is correct, functional webs

could have decay times of 10 seconds or more (Fuster, 1995, 1998b, 2000;

Zipser et al., 1993). Assuming that this feature generalizes to the proposed

word webs (see Chapter 4), the indirect sequence detectors responding toword-pair sequences may well be assumed to operate on a time scale of several seconds as well.

In summary, the proposal is that mediated sequence processing known

from other neuroscientific domains is an important mechanism for syntactic

processing. In contrast to the already known mechanisms of sequence de-

tection that operate at the single neuron level, single neurons represent the

input and mediate the sequence, the present proposal puts that the same

type of mechanism exists at the level of functional webs. Thus, the relevantsequence detectors would be functional webs responding to sequences of 

neuron populations related to the processing of single words. A sequence

detector would become active if the word Ai from a word category A isfollowed by a word B j  from category B, thereby activating the correspond-

ing functional webs αi and β j  sequentially. Frequent cooccurrence of words

in linear sequences may be an important factor for establishing neuron en-

sembles specializing in the detection of word sequences. This allows for an

economic representation of word-pair sequences, largely independent of theactual delay between the words within a sentence.In Chapters 10–12, the idea roughly sketched here is more precisely for-

mulated, extended, and illustrated by putative grammar mechanisms. The

final section of this chapter may give a first idea how this proposal might

connect to syntactic theories and how it differs from them.

9.3 Sequence Detectors and Syntactic Structure

The postulate that word sequences are assessed by sequence detectors leads

to a novel view on syntactic processes. The dominating view in linguistics

has been that a hierarchical tree of syntactic category representations is

built up to parse a sentence and that the individual words of the sentence

are attached to the tree as its leaves (see Chapter 7). The tree would have the

sentence symbol S as its root, and branches would lead to phrase nodes and

Page 181: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 181/331

164 Sequence Detectors

Figure 9.1. (a) A phrase structure repre-sentation of the sentence “He comes.”Lines represent structural relationships.Abbreviations: Ppn: personal pronoun;

 V: verb; Vs: verb suffix; Np: noun phrase; Vp: verb phrase; S: sentence. (b) Puta-tive neuronal circuit based on medi-ated sequence detection. Circles rep-resent functional webs. Labels close tocircle indicate the morphemes repre-sented by word webs (upper line) and

the sequences of lexical category mem-bers sequence detectors are assumedto be sensitive to (lower line). Thin andthick lines represent qualitatively differ-ent neuronal connections between se-quence detectors and word/morphemewebs, respectively.

to lexical category nodes. Another example syntactic tree for a maximally

simple sentence, or sentence fragment, is presented in the upper diagram

of Figure 9.1. The phrase structure of the string “He comes” is used as an

example. As illustrated in Chapter 7 with different examples, this sentence

can also be broken down into noun phrase (Np) and verb phrase (Vp). Thenoun phrase is, in this case, realized as a personal pronoun (Ppr), and the

verb phrase as verb stem (V) plus verb suffix (Vs).

The tree representation per se has the disadvantage that it cannot captureone type of long-distance dependency called agreement (see Chapter 7). The

relationship between the sentence-initial pronoun and the sentence-final suf-

fix requires an extension of the concept of a two-dimensional tree structure.As detailed in Chapter 7, linguists have proposed supplementary mecha-

nisms to model the interdependence of these elements, which, in the present

example, agree in person and number. The most popular approach proposes

that features of the items be transported through the branches of the tree to

mediate between its leaves. A disadvantage of this strategy is that it is noteconomical: Two mechanisms – tree construction and within-tree transportof features – are postulated, although one may speculate that one mechanism

could equally well represent and process the multiple relationships between

the constituents of the sentence (see Chapter 7).

A syntactic model based on sequence detectors replaces the tree construct

by a set of neuronal elements mediating between word webs (Figure 9.1,

Page 182: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 182/331

9.3 Sequence Detectors and Syntactic Structure 165

bottom diagram). Separate sequence detectors responding to word pairs,

in example (a) to the pronoun–verb sequence, (b) to the verb–verb suffix

sequence, and, in the very same way, (c) to the pronoun–verb suffix sequence,are envisaged to be activated by the word string. The activation of these threesequence detectors would also represent and process structural information

about the word string. Each sequence detector would represent and, given it

is active, store the information that a particular pair of syntactically related

morphemes has recently been present in the input.

At first glance, this approach may appear more economical than the

syntactic-tree-and-feature-transport approach because it postulates only one

unified mechanism, mediated sequence detection, which, in this case, canreplace two different mechanisms, for subordination and agreement in syn-

tactic trees.

Psychological phenomenon may provide additional support for the neu-

robiologically inspired idea that mediated sequence detection operating oncategories of words could underlie syntactic processing. An important ob-

servation is that previously perceived syntactic structures are being imitated

in subsequent verbal actions. This phenomenon attributed to a mechanism

dubbed syntactic priming occurs with above-chance probability in both con-versations and controled experiments (Bock, 1986; Bock, Loebell, & Morey,

1992; Pickering & Branigan, 1999). The phenomenon is independent of whether or not the two sentences share words. A double object sentence

as a prime (“showed the children the pictures”) yields later production of 

other double object sentences (“gave the butcher the knife”), and a simi-

lar priming effect can be observed for the prepositional object paraphrase

(“showed the pictures to the children”). It has been pointed out that certain

aspects of syntactic priming are difficult to explain on the basis of some of the available syntactic theories (for discussion, see Pickering & Branigan,

1999).

It is obvious that imitation of sequences of words from similar lexical

categories can be explained easily and straightforwardly based on the notion

of sequence detectors operating on categories of word representations. If 

sequence detectors play a role in both the detection and production of stringsof words from given lexical categories, the explanation of syntactic priming

can be based on primed sequence detectors. Priming of these neuronal websby an incoming sentence enhances the activity level of these neuronal units,

thus later enhancing the probability that similar word sequences will be

produced.

There cannot be any doubt that networks made up of neurons can realizeimportant aspects of the serial order of events. It is, nevertheless, impor-

tant to point to some of the neurocomputational research that investigated

Page 183: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 183/331

166 Sequence Detectors

in detail the mechanisms discussed here. There is a long history of work

exploring the capabilities of associative networks, which has been sparked

by theoretical proposals (e. g., McCulloch & Pitts, 1943) and empirical re-sults (e.g., Reichardt & Varju, 1959). Willwacher (1976, 1982), for exam-

ple, presented an early implementation of a single-layer network capableof learning and retrieving letter sequences, and Buonomano (2000) recently

showed that a variety of precise delays between events can be learned and

represented in an associative network consisting of excitatory and inhibitory

neurons that are organized in one neuronal layer. Some researchers have in-

cluded into their simulations much detail about the specific features of the se-

quences under study, such as complex grasping or walking movements (Cruse& Bruwer, 1987; Cruse et al., 1995), and about neuronal responses as re-vealed by neurophysiological investigation (Kleinfeld & Sompolinsky, 1988).

Apart from single-layer associative networks, more complex networks

have been used with some success. Elman used an architecture that includes

hierarchically organized layers, one of which is reciprocally connected to an

additional so-called memory layer where information about past events can

accumulate. This architecture proved particularly fruitful for modeling serial

order of language elements (Elman, 1900; see Chapter 6). Dehaene useda three-layer model, including one layer where sequence detectors could

develop very similar to the ones discussed here in the context of syntactic

processes (Dehaene, Changeux, & Nadal, 1987).

Despite these successes in modeling serial order relationships in neural

models, it should be kept in mind that the successful application of a net-work to a problem of serial order does not clearly imply that the relevant

mechanisms on which sequence production or detection is based have been

uncovered. For some simulation approaches it remains to be shown whetherthe crucial mechanism is direct sequence detection by delay lines or, as an

alternative, mediated sequence detection relying on separate neuronal unitsdevoted to the processing of serial order information. This can be decided

by looking closely at the behavior of individual neurons included in the

network.

The model of sequence detection discussed here makes specific predic-

tions on the outcome of neurophysiological experiments that have, as to

the author’s knowledge, not been carried out yet. The considerations onsyntax offered in this section would suggest that it might be advantageous

to have neuronal units available that respond specifically to a sequence of 

events A and B, but that their response is largely independent of the delay.

A further prediction might be that the relevant delays range between 0.2

seconds and tens of seconds. The model discussed here would suggest that

such sequence detectors responding to specific word sequences would be

Page 184: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 184/331

9.3 Sequence Detectors and Syntactic Structure 167

particularly common in the left perisylvian cortex. Furthermore, the neuro-

biological approach may provide a brain-based explanation of neurophysi-

ological and metabolic changes in brain activity related to the processingof syntactic information (Caplan, 2001; Friederici, Pfeifer, & Hahne, 1993;

Hagoort, Brown, & Groothusen, 1993; Indefrey et al., 2001; Moro et al.,2001; Neville et al., 1991; Osterhout & Holcomb, 1992). This issue will be

raised again in Chapter 13.

After the illustration of how sequence detectors may provide a basis for

grammar, it now becomes important to formulate the proposal in a more

general way.

Page 185: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 185/331

CHAPTER TEN

Neuronal Grammar

Large, strongly connected groups of neurons were proposed to form the

neurobiological substrate of higher cognitive processes in general and lan-

guage in particular. If the reader wishes, the ultimate answer to the question

of language, the brain, and everything was suggested to be neuronal ensem-ble. Different authors define terms such as neuron ensemble, cell assembly,

and neuronal group in different ways, and therefore a new term, functionalweb, was proposed and its meaning clarified (see Chapters 2, 5, and 8). There

is support for the concept of functional webs from neurophysiological and

neuroimaging experiments on language and memory (see Chapters 2 and 4).In this chapter, the notion of a functional web is used as a starting point for

a serial-order model. The elements of this model are called neuronal sets.

Neuronal sets are functional webs with additional special properties that are

relevant for serial-order processing. Neuronal sets can represent sequences

of words and are then called sequence sets (or alternatively, sequencing units,or sequence detectors). New terms are introduced to distinguish the entitythat has empirical support (functional web) from the theoretical concept

developed (neuronal set).

In this chapter, the notions neuronal set  and sequence set  are explained

and applied to introduce a putative basic mechanism of grammar in the

brain. Properties of functional webs are first briefly summarized, and then

the concept of a neuronal set  is defined as a special type of functional web.

The process of the detection of word sequences is spelled out in terms of theordered activation and deactivation of neuronal sets. The possible neuro-

biological correlate of lexical categories is discussed.

168

Page 186: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 186/331

10.2 Neuronal Sets 169

10.1 The Story So Far

A functional web is a widely distributed group of cortical neurons that arestrongly connected to each other (see Chapter 2). Because of its strong in-

ternal connections, the web is assumed to act as a functional unit, that is,

stimulation of a substantial fraction of its neurons lead to spreading activity

and finally to full activation of the set. This full activation is called ignition.Although after an ignition has taken place refractory periods of neurons and

other fatigue effects again reduce the excitation level of the set, activity in the

web is proposed not to die out completely. The strong internal connections

still allow activity to be retained in the web. A wave of activity may rever-berate in the set, as proposed by Hebb (see Chapter 8). Results of studies of 

cognitive processes using noninvasive physiological recordings are consis-tent with the view that ignition and reverberation are successive processes

reflected in stimulus-evoked cortical potentials and induced high-frequency

activity (see Chapter 4).

It is important to note the differences between the concepts of rever-

beration and ignition. Ignition is a brief event, whereas reverberation is a

continuous process lasting for several seconds or longer. Ignition of a func-tional web involves all of its neurons, or at least a substantial proportion of 

them, whereas reverberation can be maintained by small neuron subgroups

within the set being active at given points in time. Ignition does not imply

a fixed spatiotemporal order of neuronal activity. In contrast, reverberation

is characterized by a fixed sequence of neuron activations, a defined spatio-

temporal pattern of activity within functional webs. Ignition would be theresult of the overall strong connections within a web, and reverberation

would be made possible by its strongest internal links envisaged to provide apreferred highway for spreading neuronal activity, so to speak. Importantly,

two distinct types of activation of a web are postulated. In addition, there

is the possibility that the web is in an inactive state – that is, that it stays at

its resting level. Therefore, a functional web can have three activity states:It can rest, reverberate, or ignite.

10.2 Neuronal Sets

Neuronal sets are functional webs characterized by a greater variety of ac-tivity states.

If different functional webs are connected to each other, the excitatory

processes of ignition and reverberation in a given ensemble likely influence

other webs directly connected to it. Through connections between webs,

which are assumed to be much weaker, in the average, than their internal

Page 187: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 187/331

170 Neuronal Grammar

connections, one set of neurons can have an activating effect on the other.

This between-web stimulation does not necessarily mean that there is

a long lasting effect in the second set. A continuously reverberating andautonomous wave of activity is not necessarily generated in the stimu-

lated set, and an ignition only takes place if the activation from outsideis substantial, above an ignition threshold. Subthreshold activation of a web

as a result of input from one or more other sets is called priming at the

neuronal level . Neuronal sets are conceptualized as functional webs that can

exhibit priming. Thus, neuronal sets are characterized by four different types

of possible activity states: ignition ( I ), reverberation (R), priming (P ), and

inactivity (0).To clarify the main differences between terms used in this book to refer

to neuron groups of different kinds, Table 10.1 gives an overview. The

terms neuron group, cell assembly, functional web, neuronal set , and, finally,  sequence set  or web refer to subcategories with increasingly narrower

definitions.

The term priming must not be mixed up with priming at the behaviorallevel (e.g., semantic priming; see Chapter 5), although it is possible that the

neuronal mechanism is the cause of the behavioral phenomena (Milner,1957). Because later chapters deal with neuronal mechanisms, the term priming usually refers to the neuronal level unless otherwise indicated.

For a network formed by many neuronal sets, with stronger and weaker

connections between the sets, it appears plausible to assume that differentactivity states of one set may differentially affect other connected sets. Be-

cause ignition is a substantial excitatory process, it may affect other directly

connected sets, or neighbor sets, regardless of how strong connections are

between the sets. In contrast, the less powerful process of reverberation, towhich less neurons contribute at any point in time, may be visible in neigh-

bor sets only if between-set connections are strong. Ignition is assumed toprime connected neighbor sets, whereas reverberation is assumed to prime

neighbor sets only if the connection between the sets is strong.

In the state of reverberation, a neuronal set is similar to a McCulloch–Pitts

neuron of threshold one with one self-connecting loop. However, as already

mentioned in Chapter 2 and later, neurophysiological evidence does not

support the notion of a neuronal element exhibiting constantly enhancedactivity over longer periods of time (Fuster, 1995). Instead, activity levels

of memory cells in many cases exhibit an exponential decrease of activity.

If the memory cells reflect the activity state of larger sets, it makes sense

to assume that also the larger sets show exponential decline of activity with

time if the set is in its reverberatory state. The activity level A of a neuronal

Page 188: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 188/331

       T     a       b       l     e

       1       0  .

       1  .     O    v    e    r    v     i    e    w

    o      f    t     h    e     d     i      f      f    e    r    e    n    t    t

    e    r    m    s    u    s    e     d    t    o    r    e      f    e    r    t    o     d     i      f      f    e    r    e    n    t    t    y    p    e    s    o      f    n    e    u    r    o    n    g    r    o    u    p    s .     T    e    r    m    s ,    t     h    e    m    o    s    t    c    r    u    c     i    a     l

      f    e    a    t    u    r    e    s    o      f    t     h    e

     i    r    r    e      f    e    r    e    n    t    s ,    a    n     d    a    s    p    e    c    t    s

    o      f    t     h    e     i    r    u    n     d    e    r     l    y     i    n    g    p    u    t    a    t     i    v    e    m    e    c     h    a    n     i    s    m    s    a    r    e    s    u    m

    m    a    r     i    z    e     d .     (     F    o    r    e    x    p     l    a    n    a    t     i    o

    n ,

    s    e    e    t    e    x    t .     )

     C   o   n   c   e   p    t   s   o     f     C   e     l     l     A   s   s   e   m     b     l     i   e   s

     T   e   r   m

     M   e   a   n     i   n   g

     M   e   c     h   a   n     i   s   m

     N    e   u    r    o    n    g    r    o   u    p

     S    e     l    e    c    t     i    o    n    o     f    n    e   u    r    o    n    s     (    m    a   y    n    o    t     b    e    c    o    n    n    e    c    t    e     d     )

     N    e   u    r    o    n    a     l    e    n    s    e    m

     b     l    e

     S    e     l    e    c    t     i    o    n    o     f    n    e   u    r    o    n    s    t     h    a    t

     N    e   u    r    a     l    a    s    s    e    m

     b     l   y

     C    e     l     l    a    s    s    e    m     b     l   y

     A    r    e    s    t    r    o    n    g     l   y

    c    o    n    n    e    c    t    e     d    t    o    e    a    c     h    o    t     h    e    r

     A    c    t    a    s    a     f   u    n    c

    t     i    o    n    a     l   u    n     i    t ,

     i    g    n     i    t    e

     A    s    s    o    c     i    a    t     i   v    e     l    e    a    r    n

     i    n    g     i    n    a   u    t    o    a    s    s    o    c     i    a    t     i   v    e    m    e    m

    o    r   y

     S    t    r    o    n    g    c    o    n    n    e    c    t     i    o

    n    s     b    e    t   w    e    e    n    n    e   u    r    o    n    s

     F   u    n    c    t     i    o    n    a     l   w    e     b

     S    e     l    e    c    t     i    o    n    o     f    n    e   u    r    o    n    s    t     h    a    t

     A    r    e    s    t    r    o    n    g     l   y

    c    o    n    n    e    c    t    e     d    t    o    e    a    c     h    o    t     h    e    r

     A    c    t    a    s    a     f   u    n    c

    t     i    o    n    a     l   u    n     i    t ,

     i    g    n     i    t    e

     M    a     i    n    t    a     i    n    a    c    t     i

   v     i    t   y ,    r    e   v    e    r     b    e    r    a    t    e

     A    s    s    o    c     i    a    t     i   v    e     l    e    a    r    n

     i    n    g     i    n    a   u    t    o    a    s    s    o    c     i    a    t     i   v    e    m    e    m

    o    r   y

     S    t    r    o    n    g    c    o    n    n    e    c    t     i    o

    n    s     b    e    t   w    e    e    n    n    e   u    r    o    n    s

     R    e   v    e    r     b    e    r    a    t     i    n    g    s   y

    n     fi    r    e    c     h    a     i    n    s    c    o    n    n    e    c    t     i    n    g    n    e

   u    r    o    n    s

     N    e   u    r    o    n    a     l    s    e    t

     S    e     l    e    c    t     i    o    n    o     f    n    e   u    r    o    n    s    t     h    a    t

     A    r    e    s    t    r    o    n    g     l   y

    c    o    n    n    e    c    t    e     d    t    o    e    a    c     h    o    t     h    e    r

     A    c    t    a    s    a     f   u    n    c

    t     i    o    n    a     l   u    n     i    t ,

     i    g    n     i    t    e

     M    a     i    n    t    a     i    n    a    c    t     i   v     i    t   y ,    r    e   v    e    r     b    e    r    a    t    e

     C    a    n     b    e    p    r     i    m    e     d

     A    s    s    o    c     i    a    t     i   v    e     l    e    a    r    n

     i    n    g     i    n    a   u    t    o    a    s    s    o    c     i    a    t     i   v    e    m    e    m

    o    r   y

     S    t    r    o    n    g    c    o    n    n    e    c    t     i    o

    n    s     b    e    t   w    e    e    n    n    e   u    r    o    n    s

     R    e   v    e    r     b    e    r    a    t     i    n    g    s   y

    n     fi    r    e    c     h    a     i    n    s    c    o    n    n    e    c    t     i    n    g    n    e

   u    r    o    n    s

     E     f     f    e    r    e    n    t    c    o    n    n    e    c    t     i    o    n    s    t    o    t     h    e    n    e   u    r    o    n    s

     S    e    q   u    e    n    c    e    s    e    t

     S    e     l    e    c    t     i    o    n    o     f    n    e   u    r    o    n    s    t     h    a    t

     S    e    q   u    e    n    c    e     d    e    t    e    c    t    o    r

     S    e    q   u    e    n    c    e   w    e

     b

     A    r    e    s    t    r    o    n    g     l   y

    c    o    n    n    e    c    t    e     d    t    o    e    a    c     h    o    t     h    e    r

     A    c    t    a    s    a     f   u    n    c    t     i    o    n    a     l   u    n     i    t ,

     i    g    n     i    t    e

     M    a     i    n    t    a     i    n    a    c    t     i

   v     i    t   y ,    r    e   v    e    r     b    e    r    a    t    e

     C    a    n     b    e    p    r     i    m    e     d

     R    e    s    p    o    n     d    s    p    e

    c     i     fi    c    a     l     l   y    t    o    s    e    q   u    e    n    c    e    s

     A    s    s    o    c     i    a    t     i   v    e     l    e    a    r    n

     i    n    g     i    n    a   u    t    o    a    s    s    o    c     i    a    t     i   v    e    m    e    m

    o    r   y

     S    t    r    o    n    g    c    o    n    n    e    c    t     i    o

    n    s     b    e    t   w    e    e    n    n    e   u    r    o    n    s

     R    e   v    e    r     b    e    r    a    t     i    n    g    s   y

    n     fi    r    e    c     h    a     i    n    s    c    o    n    n    e    c    t     i    n    g    n    e

   u    r    o    n    s

     E     f     f    e    r    e    n    t    c    o    n    n    e    c    t     i    o    n    s    t    o    t     h    e    n    e   u    r    o    n    s

     I    n     d     i    r    e    c    t    s    e    q   u    e    n    c    e    p    r    o    c    e    s    s     i    n    g

171

Page 189: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 189/331

172 Neuronal Grammar

set Si at time t may be described by Eq. (1).

 A(Si, t ) = A(Si, t 0) e−ct 

= A(Si, t 0) e−c(t −t 0) (1)

where t 0 is the point in time when the set has received excitatory input from

neurons outside the set, t is the time difference between t and t 0, A(Si, t 0) isthe amount of activity it exhibited at t 0, and c is a constant determining the

slope of the exponential decrease of activity (c can be assumed to be 1 here).

The larger t , the smaller the exponential term e−ct . In essence, when

reverberating, the neuronal set continuously loses activity, and this activitydecline is assumed to be exponential.

As in many network simulations (see Chapter 6), the time continuum issliced into discrete time steps. This is done for illustration and simulation pur-

poses. In reality, the exponential decrease of activity is probably continuous

rather than in discrete steps.

If a set receives excitatory input from outside, an ignition may result. If 

the input is not sufficient for reaching the ignition threshold, the set is merely

being primed. In contrast to ignition and reverberation, priming of a givenset depends on input from other sets. Priming therefore terminates as soon

as the input has ceased.

One may question this assumption, because excitatory postsynaptic po-

tentials are known to last for tens of milliseconds after an action potential

has arrived (see Chapter 2). However, activity dynamics of neuronal setsinvolved in processing serial order of words would probably operate at a

much longer time scale (see Chapter 8). If the time steps considered in a

model of sequence detection are on the order of 30 or so milliseconds, ex-citatory postsynaptic potentials can probably be considered as momentary

phenomena.

Again, the momentary process of priming is distinguished from the long-lasting process of reverberation. Certain forms of behavioral priming

(e.g., repetition priming) would therefore need to be attributed to rever-

beration rather than to priming at the neuronal level. From a neurobio-

logical point of view, the concept of priming as it is used in psychology

may correspond to a variety of different neuronal mechanisms (see Mohr &Pulvermuller, in press, for further elaboration).

Under priming, the various excitatory inputs to a given set from its ac-

tive neighbors summates at each time step. Inputs from neighbor sets that

reverberate or ignite are being received at the same time through different

input fibers of the neurons of the set, and this can lead to spatial summation

in its neurons. One may also say that the neuronal set as a whole summates

Page 190: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 190/331

10.2 Neuronal Sets 173

the various simultaneous inputs so that its overall activity level increases.

If additivity of activity is assumed, the following equation gives the activity

level of a setSi at time

t , which received input from

j other sets

S j  one timestep before:

 A(Si, t ) =  j A(S j , t − 1) (2)

where A(S j , t − 1) indicates the amount of excitatory stimulation a set S j sends to set Si at a given time step t − 1 and which arrives at Si at t . Although

Eq. (2) may give a rough idea about the real processes, these are likely morecomplex because the underlying assumption of additivity is idealized.

If an already reverberating set receives stimulation from outside, the ex-citatory effects of the reverberation itself and the inputs from other sets are

again assumed to summate.

 A(Si, t ) = A(Si, t 0)−c(t −t 0) +  j A(S j , t − 1) (3)

This formula is one way to express that two main sources of excitation of 

a neuronal set are (a) its internally reverberating excitation and (b) the

input to the set through afferent connections from other neurons – that is,

through priming. As an alternative to mere summation, the activity providedby reverberation and priming may interact in a more complex manner.

As in the artificial neuron-like machines proposed by McCulloch and Pitts,

the integrated activity can be used as a criterion for the decision about fully

activating a neuronal set. If the activity level summed over space and timeexceeds a certain value, the ignition threshold θ , then the set ignites. After

its ignition ( I ), and everything else being equal, the set starts to reverberate

again at a given reverberation level, R1. To state this briefly:

 A(Si, t ) > θ ⇒ A(Si, t + 1) = I ⇒ A(Si, t + 2) = R1 (4)

The double arrows denote causality and temporal succession. One time step

after the threshold is reached, an ignition occurs which is, again after one

time step, followed by the reverberation process.

In essence, neuronal sets are assumed to have qualitatively different possi-ble activity states. At each point in time, a set is characterized by one activity

state from the set AS of possible activity state types:

 AS = {0, P , R, I } (5)

Again, 0 is the resting level; I  is full activation or ignition; R represents

reverberation, which follows ignition; andP is priming caused by the ignition

or reverberation of a different but connected neuronal set.

Whereas 0 and I  refer to single activity states, P  and R denote sets

of possible states. As Eqs. (1) and (2) imply, there are different levels of 

Page 191: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 191/331

174 Neuronal Grammar

reverberation and priming. The set R of reverberatory states is defined by

the exponential function Eq. (1) and, therefore, there is, in principle, an

unlimited number of possible activity states.R = {R1, . . . , Ri , . . . , Rn}; R1 < θ , Rn > 0 (6)

There is no upper limit for the number n of the possible reverberatory statesRi because decline of activity is continuous, and Ri is therefore from the set

of positive real numbers. Thus, theoretically speaking, there are uncountably

many possible states Ri . From this, it follows that a neuronal set cannot be

considered a finite state device because a finite state device is defined as

having only a finite number of possible states.Nevertheless, a set of neurons with infinitely many possible activity statesis unlikely to exist in reality. The noise available in all biological systems

cancels the minor differences between similar states so that, in fact, it is only

possible to distinguish a finite set of states. The influence of the noise depends

on its ratio to the signal. If a neuronal set consisted of a very large number

of neurons, each stimulated by statistically independent noise, it would be

possible to minimize or even cancel the effect of random noise to each in-

dividual neuron by averaging activity over all elements of the set at eachpoint in time (see Chapter 2). Given good signal-to-noise-ratios are being

computed, the number of possible activity states can therefore be assumed

to be high, although not infinite. The number of distinguishable states of 

reverberation of a set is limited by the signal-to-noise ratio of the system.

A similar point as the one made above for reverberation can be made forlevels of priming as well. As implied by Eq. (2), there is no principled upper

limit for the number of members of the set P of states of priming.

P = {P 1, . . . , P i , . . . , P n}; P 1 < θ, Rn > 0 (6)

Although the number of inputs to a neuronal set is limited, n can be high

because each input can be strong or weak, depending on connection strengthand the activity level in the stimulating set. Because this activity level of the

stimulating sets can vary, the input to the primed set varies as well.

Although there are theoretical and empirical reasons to conceptualize

neuronal sets as networks with properties as described by Eq. (1) to Eq. (5),

certain simplifications appear necessary to illustrate the functioning of gram-mars composed of neuronal sets. In this and the following two chapters andthe related Excursuses, grammar circuits are introduced. These are kept as

simple as possible. The circuits and their functions are designed to provide

comprehensive examples of the computational perspectives of grammars

formulated in terms of neurons. The following simplifications are made to

allow for straightforward and well-intelligible illustration:

Page 192: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 192/331

10.3 Threshold Control 175

(1) Only a few states of reverberation (labeled R1, R2, R3, etc., or, al-

ternatively, R, R, R, etc.) and priming ( P 1, P 2, . . . , or P , P , . . . ) are

mentioned explicitly in the examples that follow.(2) Continuous activity decrease with time [Eq. (1)] is ignored except if anew level of activity is being introduced. R1 denotes the highest ac-

tivity level in a network at a given point in time (irrespective of its

absolute value); R2 denotes the second highest activity level, just one

step below; R3 refers to the third-highest levels; and so on. The same

applies for levels of priming.(3) Spatiotemporal summation of activity and the criterion for ignition

is simplified.

Again, this is done to maximize intelligibility of the formalisms and make

computations less complex and therefore easier to illustrate. The exponential

activitydeclineofneuronalsetswithtimeisproposedtobecrucialforstorage

of serial-order information.

10.3 Threshold Control

In Chapter 5, it was argued that for a cortex-like network to operate prop-

erly, it is necessary to assume a regulation mechanism that keeps the activity

level within certain bounds. A feedback regulation mechanism can be as-

sumed to adjust the thresholds of individual neurons by way of providing or

removing background activity to move the cortical activity level close to atarget value. This mechanism has been called the threshold control mecha-

nism (Braitenberg, 1978a; Braitenberg & Schuz, 1998; cf. Chapter 5).

Here, it is assumed that the regulation mechanism detects fast activity in-creases and removes background activity to the cortex if a substantial activity

increase is detected. If the level of activity is generally low, the mechanism

provides additional excitation. As emphasized in Chapter 5, other criteriafor activation of the regulation mechanism can easily be imagined, and at

this point, it is not clear which of the several possibilities is actually realized

in the real brain. However, because it is evident that a feedback regulation

mechanism of some kind is necessary, one of the several options must be

chosen for realistic modeling.A criterion based on activity increase can be realized in a network without

difficulty. The global activity state summed over all neurons in the network

at a time point t  could be compared to global activity at an earlier time

(e.g., one time step earlier, at t − 1). Only if the difference exceeds a certain

value is inhibitory input to the network provided. An even simpler crite-

rion for activity increase can be introduced on the basis of the following

Page 193: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 193/331

176 Neuronal Grammar

consideration. The most extreme activity increase is probably present in a

network of neuronal sets if one or more sets ignite after being in the rest-

ing state 0. Therefore, such fast activity increase may be assumed to causethe regulation mechanism to become active and initiate a process of global

inhibition affecting the entire network.Compared to the ignition of an inactive set, the activity change is smaller

if an already primed or reverberating neuronal set ignites. Ignition implies

simultaneous activity of the majority of the neurons in the set. The num-

ber of neurons in a preactivated set that become active during an ignition

(in addition to those already being active) is therefore smaller compared

to a previously inactive set that ignites. It is assumed that full ignition of a neuronal set at rest activates the threshold control mechanism, whereas

the smaller activity increase induced by the primed ignition of an already

preactivated or primed set does not. The symbol I ∧ is used to refer to a full

ignition that subsequently activates threshold regulation, whereas I denotes

ignition without any consequences for the threshold regulation machinery.

The expression preactivity of a set  is used to refer to any activity statedifferent from rest (0) and ignition ( I ); that is, either reverberation (R) or

priming (P ).A full ignition I ∧ leads to global inhibition in the network. This is realized

as follows: When I ∧ occurs in one set, all preactivity states of other sets – that

is, reverberation and priming levels – is diminished. R1 and P 1 are changed

to R2 and P 2, and, generally, all states n are changed to n+ 1. Recall that the

numbergivestherankinthehierarchyofactivitystates,1beinghigherthan2.

A full ignition I ∧ reducing preactivity states therefore changes all states Rn

and P n to Rn+1 and P n+1, respectively.

If inhibition happens regularly, each reverberating set loses activity overtime. Thus, a threshold regulation mechanism becoming effective with con-stant frequency may cause constant activity decrease, as described by Eq. (1),

in preactivated neuronal sets. Furthermore, a global inhibitory mechanism

could guarantee that the slopes of activity decline are the same in different

sets, an assumption relevant for the pushdown memory mechanism focused

on in Chapter 12 and Excursus E5 (see also Pulvermuller, 1993).

As mentioned, a threshold regulation mechanism can also have an excita-

tory effect. It is assumed that if activity levels in all reverberating and primedneuronal sets are generally low, additional global excitation is provided by

the regulation mechanism. All preactivity levels therefore become higher.

Because the formulation of preactivity states refers to relative activity

levels (e.g., “R1” meaning “best activated set,” not a particular absolute

activity level), the activating function of the regulation mechanism is builtinto the formulation, so to speak. One or more sets is always at the highest

Page 194: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 194/331

10.4 Sequence Detection in Networks of Neuronal Sets 177

level of reverberation and priming. A global increase of activity shows up

only as relabeling of activity levels if there is no neuronal set in the network

exhibiting the highest levels of reverberation and priming (R

1 andP 

1). Inthis case, the numbering is adjusted so that the highest activity level is 1.

10.4 Sequence Detection in Networks of Neuronal Sets

Based on which wiring would a neuronal unit respond to a sequence AB

of events but not to BA? As outlined in Chapter 9, one could have strong

connections from an input unitα, specifically responding to the occurrence of 

an A in the input, to a sequence detector, and between the sequence detectorand β, the neuronal unit responding to the occurrence of a B in the input.

Through these strong bottom-up connections, information about activityof both α and β could influence activity of the neuronal set γ  mediating

sequence detection.

A few terminological issues should be mentioned at this point. Neuronal

sets that respond specifically to the ordered sequence of activations of two

other sets are called sequence sets. The words sequence detector and sequenc-

ing unit are also used to refer to neuronal sets mediating sequence detection.If the sequence set γ  only responds to the event “first α, then β,” but not to

the reverse sequence, the event A (to which α responds) is called the earlier 

event and the event B (to which β responds) is called the later event .

The issue of mediated sequence detection has already been addressed in

the context of neural networks (Chapter 6) and brain mechanisms of serial or-

der (see Chapter 9). As pointed out, there are different mechanisms availablefor realizing sequence detection. In McCulloch–Pitts networks, sequence

sensitivity is established by connections, from event detectors α and β tothe sequence detector γ , with different conduction delays. Delay neurons

were introduced to guarantee the delay (see Section 6.1). The Reichardt–

Varju theory proposed low–pass filtering of the signal from one of the eventdetectors (responding to the earlier event) as a possible mechanism crucial

for sequence detection (Egelhaaf, Borst, & Reichardt, 1989; Reichardt &

Varju, 1959). In this case, the delay of the earlier signal would be provided

by a mathematical transformation carried out by nerve cells.

A further possibility is that sequence sensitivity is merely the result of connection strength between event and sequence detectors, and to the act-ivity dynamics of the former. Assume, for example, that the event detectorβ responding to the later event B has stronger connections to the sequence

detector γ  than α which responds to the earlier event A. Because of the

strongerβγ connection, any activity inβ excitesγ more strongly than activity

in α. Assume further thatα and β are characterized by the same exponential

Page 195: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 195/331

178 Neuronal Grammar

decline function of reverberating activity [Eq. (1)]. Now, if A occurs before B,

the declined small signal through the weakαγ  connection summates with the

full strong signal through theβγ 

connection. This sum is greater than theβγ  input alone. If B occurs before A, the declined strong signal throughthe βγ  connection summates with the not degraded weak signal through

the αγ  connection. The latter sum is smaller than the βγ  signal as soon

as the degradation of the strong input through βγ  is larger than the full

weak input through αγ . Thus, if the decline function is steep compared to

the difference between the strong and the weak full signals, the summed

input to γ  is greater if A occurs before B than in response to the reverse

sequence of events. Direct connections from both α and β to γ  are sufficientfor providing γ  with serial-order information. The threshold for activatingγ  can be adjusted so that it responds to the activation sequence “first α, thenβ,” but not to the reverse sequence.

In summary, different mechanisms may mediate the sensitivity to serial-order information of neurons and sets of neurons. Because the issue of which

of the alternative mechanisms underlie grammar processing in the human

brain is still far from being addressed directly by empirical research, it suffices

to postulate a mechanism of mediated serial-order detection carried out atthe level of neuronal sets (see Chapter 9). The exact details of how this

mechanism is realized are left for future research.Information flow from event to sequence detectors is required for me-

diated serial-order detection. This could be implemented by unidirectional

connections from word webs to sequence sets. However, bidirectional con-

nections between neuronal sets is always assumed to be the default in the

present context. The motivation for this is two-fold: One reason comes from

neuroanatomical and neurophysiological studies (see Chapters 2 and 8).If long neuronal connections in the cortex are considered the substrate

of grammar-relevant mechanisms, one-directional connections are proba-

bly the exception rather than the rule. If two cortical areas are connected

to each other in one direction, they usually exhibit the reciprocal connec-

tion as well (Pandya & Yeterian, 1985; Young et al., 1995). At the level of 

local cortical circuits, Abeles and colleagues (1993) found evidence for re-verberatory loops, a fact implying bidirectional rather than unidirectional

links between many of the neurons in the circuit. Because neuroanatomicalstudies indicate that most connections, local and between areas, are recipro-

cal, unidirectional connections between large distributed neuronal sets are

less likely (see the reciprocicity principle in Chapter 2). The default assump-

tion should be reciprocal connections. Still, the reciprocal connections maybe asymmetrical; that is, stronger in one direction than in the other. This is

taken into account later in this chapter.

Page 196: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 196/331

10.4 Sequence Detection in Networks of Neuronal Sets 179

The second reason for assuming bidirectional links is psycholinguistic

in nature: It appears that when listening to a sentence, the first words

usually lead one to generate hypotheses about the next. In some cases,however, words occurring later in a sentence help disambiguate or fully

understand earlier sentence parts. The verb in “Betty switched” can be un-derstood in different ways, and it is only after perception of later sentence

parts that the appropriate reading can be determined. The final part of the

sentence, “her shopping mall,” would lead to one interpretation of the verb,

whereas if the final part of the sentence is “her TV off” would elicit another.

This can be modeled by backward connections from sequence sets to the

word webs.For these reasons, neuroscientific and psychological in nature, one mayprefer having activity spread in both directions, back and forth, through the

representation of a sentence being heard, parsed, and understood. Recip-

rocal connections are therefore assumed to connect word webs to sequence

sets and sequence sets among each other.

The reciprocal connections between word webs and sequence sets is

assumed to be asymmetrical. The proposed connection in one direction

would be stronger than the connection in the other. If two sets are connectedreciprocally but one of the sets usually becomes active before the other, it

may be that the connections from the first to the second set become stronger

than the links back from the second set to the first. This is an unavoidable

conclusion if the following assumptions are being made: (a) Hebbian cor-

relation learning takes place, that is, synaptic strengthening is strongest if simultaneous pre- and postsynaptic activity occurs; (b) all conduction times

are non-zero; (c) some conduction delays equal frequent delays between

ignitions of two sets (Miller, 1996). Together, these assumptions motivate theassumption of asymmetry in reciprocal connections between neuronal sets.

In this view, information about possible input sequences is represented

in a network (i) by links between neuronal sets and (ii) by the direction inwhich connections are stronger, assuming that the connection is strong in

one direction and weak in the other.

Distinguishing only between “strong” and “weak” connections is certainly

another simplification. Strength of synaptic connections in the brain can

take a large number of different values. However, this minimal distinct-ion between connection strengths allows for complex syntactic processing

in neuronal grammar, as exemplified later. The additional advantages of 

assuming more degrees of freedom for connection strengths is mentioned

occasionally in this text, but is not explored systematically.

Furthermore, the assumption that all reciprocal connections between

word webs and sequence sets are asymmetric may turn out to be too strong.

Page 197: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 197/331

180 Neuronal Grammar

Figure 10.1. Two input units or word webs connected to a se-quence set. All circles represent neuronal sets that can be primed,can ignite and reverberate after their ignition. (Top) Strong con-nections link input unit α to the sequence detector, and the se-quence detector in turn to set β. Weak connections run in theopposite direction. The sequence detector becomes active if itis stimulated twice at a time.

In addition, there may be neuronal sets that are sensitive to both possi-

ble sequences of two events, AB and BA. These neuronal sets are realized

as reciprocal symmetric connections to their input units. This kind of Anddetector may make it easier to model languages in which there is no preferredorder of constituents.

The upper diagram in Figure 10.1 presents a sequence set connected

with two input units or word webs. This network is assumed to be capa-

ble of sequence detection, that is, it exhibits a certain activity pattern that

can be taken as an indicator that a string in the input is consistent with

the network structure. In linguistic terms, this may be similar to the judg-

ment that the string in the input is grammatical or well formed. This net-work differs from those discussed in earlier chapters because all circles in

Figure 10.1 denote neuronal sets that can reverberate, ignite, and be primed

by other sets. In earlier diagrams, circles represented neurons, local clusers of 

neurons, or cell assemblies with less sophisticated activity dynamics

(see Table 10.1).In the upper diagram of Figure 10.1, thick and thin arrows are used to in-

dicate relatively strong or weak connections, respectively. The weaker con-

nections make the diagrams more complex, in particular if numerous setsand connections are part of it. The weak connections are therefore omit-

ted, as in the lower diagram of Figure 10.1, which is supposed to have the

same meaning as the upper diagram. The same difference exists between

the upper and lower diagrams in Figure 10.2. In the diagrams at the bottom,and in all subsequent figures, arrows indicate that there is a strong connec-

tion in the direction of the arrow and a weaker connection in the reverse

direction. Double arrows may indicate symmetrical connections.

Page 198: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 198/331

10.5 Activity Dynamics of Sequence Detection 181

Figure 10.2. Input units connected through twosequencing units: a forward-looking unit αf  and abackward-looking unit βp. Sequencing units are du-plicated to make formulation of neuronal grammarseasier.

In summary, sequence sets may specifically respond to serially orderedignitions of two neuronal sets corresponding to words (word webs). The

mechanism of mediated sequence detection requires information flow, and

therefore connections, from input units (e.g., word webs) to sequence detect-

ors (e.g., sequence sets). The sequence detecting set and the event detectors

are assumed to be linked reciprocally to each other; however, with asym-metrical weights. Stronger connections are assumed in the direction of the

usual activity flow, from the event detector of the earlier event to that of 

the later event, compared to the respective links in the opposite direction.Although it is clear that the mechanism sketched here is in need of both fur-

ther theoretical elaboration and empirical support, it nevertheless becomes

possible to explore its function in an envisaged grammar network.

10.5 Activity Dynamics of Sequence Detection

The basic idea put forward here is that word webs and the sequence setsconnected to them can respond in a specific manner to a sequence of words

and fail to do so in response to the same words presented in a different order.The mechanism this is grounded in is the response of the sequence set to

sequences it is prepared to detect. If the sequence set γ  is connected to both

the early event’s web α and to the late event’s web β, and if γ  needs two

inputs to become active, ignition of α may first prime γ , so that the ignition

of β can later fully activate γ . The primed ignition of the sequence set would

Page 199: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 199/331

182 Neuronal Grammar

then be the neuronal indicator of the fact that the input was in accord with

the network structure.

Before examining how a network composed of word webs and sequencesets would operate, it is necessary to introduce a few assumptions about

activity dynamics within neuronal sets and activity exchange caused by con-nections between neuronal sets.

Each word or morpheme in the input activates its input unit or word web

that, as a consequence, ignites and thereafter stays in the state of reverberat-

ion. As already mentioned, ignition and reverberation of one neuronal set

are assumed to have different priming effects on other sets connected di-

rectly to the igniting or reverberating circuit. Reverberation causes primingonly through strong connections. Ignition causes activity changes through

all connections, strong and weak.

Ignitions are assumed to spread from input units to sequence sets, but

only to preactivated ones. Thus, to activate a sequence set, it first needs to be

primed, or be in the reverberatory state, and then a second strong activating

input caused by an ignition must reach it. The latter assumption is somewhatsimilar to the assumption of a threshold of two of McCulloch–Pitts neurons.

Summation of activity already present in a web and the afferent input to ithas been specified by Eq. (3). Clearly, if preactivity is low, the second input to

the sequence set may not cause it to reach the ignition threshold. Therefore,

ignitions are allowed to spread only to a sequence set if it reverberates or

is being primed at high levels of  R or P . These assumptions are elaboratedfurther later.

The reader is invited to consider how the network in Figure 10.1 operates.

To this end, the changes of activity states in the course of a sequence of inputs

are illustrated.Supposethe sequence AB is in accord with the network structure, whereas

BA is not. An input sequence in accord with a network structure is calleda congruent input , whereas a sequence without corresponding wiring in the

network, and not otherwise represented, is called an incongruent input . This

is analogous to the linguistic classification of strings as grammatical, or well

formed, vs. ungrammatical, or ill formed. In contrast to the linguistic distinc-

tion, which relies on judgments of native speakers of a language, the terms

congruent and incongruent are defined on the basis of putative neuronal cir-cuits. In the circuit in Figure 10.1, the following processes may occur and

contribute to sequence detection.

Network response to a congruent input AB is as follows:

(1) The word A occurs in the input, and the word web (or input unit)α ignites.

Page 200: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 200/331

10.5 Activity Dynamics of Sequence Detection 183

(2) Because α was not primed and thus exhibits unprimed or full ignit-

ion I ∧, the threshold regulation mechanism is activated. However,

at this point, there is no other activated set, and therefore thresholdregulation has no effect. All other sets stay inactive.(3) The ignition of  α temporarily primes its connected sequence de-

tector and input unit. This does not cause an ignition of the se-

quence set because it had been inactive and would need two inputs to

ignite.

(4) After its ignition, the activated set α reverberates at the highest re-

verberation level R1. For some time, its reverberation primes the

sequence set that receives strong connections from α continuously.Because there are also strong connections from this sequence set to

the other word web, β, that represents a possible successor word B,

this set β is primed, too. At this point, one input unit, α, reverberates,

and both the sequence detector and the second input unitβ are beingprimed.

(5) B occurs in the input and the primed input unit β therefore ignites.

Thisisnowaprimedignitionbecauseβ had continuously been primed

before as a result of α’s ignition and reverberation and the strongconnections from α to β.

(6) The ignition of β is communicated to the sequence detector. Becauseit is still being primed by the reverberating α, it responds with an

ignition as well.

(7) Finally, ignition spreads back to the reverberating input unit α.

These would be the processes envisaged to occur when the network is

presented with a congruent “grammatical” string AB. The important pointis that the strong left to right connections in Figure 10.1 guarantee preactivat-

ion of the second input unit. This makes its ignition a primed ignition and,therefore, the threshold regulation machinery is not invoked. Ignitions can

finally spread backward throughout the activated network, and the primed

sequence set also ignites.

Next, the behavior of the network when an “ungrammatical” string BA

is being presented is considered.

Network response to an incongruent input BA is as follows:

(1) The word B in the input causes β to ignite.

(2) This is a full ignition I ∧ without effect on the otherwise inactive net-

work.

(3) The ignition causes temporary priming of the sequence set as a result

of backward spreading activity. Assuming that the ignition is a brief 

Page 201: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 201/331

184 Neuronal Grammar

event and backward spreading of activity is faster than the pace of 

the input elements, this temporary activation vanishes before another

input unit is activated.(4) The word web β’s ignition terminates, and this set reverberates at the

highest reverberation level R1. The sequence set and the other inputunit, α, are not being primed, because the connections they receive

from β are weak.

(5) The word A occurring in the input causes an unprimed full ignition

 I ∧ of α. This means that the activity state of α changes from 0 to  I .

(6) The strong increase in activity in the net caused by the full ignition

activates the threshold control mechanism. The level of reverberationof the already active unitβ is therefore reduced from the highest level,

R1, to a lower level, R2.

After full ignition of α, it is, in this case, not possible that ignitions spread

to the sequence detector, because it is not in the primed or reverberating

state when α ignites. The three neuronal sets are left at different states of 

reverberation and priming. Set α reverberates at R1 and set β at R2; the

sequence detector is primed by α and thus exhibits the highest priming levelP 1. No terminal wave of ignitions spreading through the network occurs, and

the set specialized for sequence detection fails to ignite.

Note, again, the main differences between the two examples: In the case

of the congruent string, there is priming of sequencing and input units re-

presenting possible successors. Such priming is not present when the deviantstring is processed as a result of the direction of strong connections. Preac-

tivation of the second input unit leads to its primed ignition (instead of full

ignition), which does not cause the threshold regulation mechanism to be-come active. Finally, processing was followed by a wave of ignitions running

from β to the sequence set and back to α, leaving all neuronal sets in the

highest state R1 of reverberation. In contrast, the backward wave of activityis not created by the ill-formed string, and the preactivated input units finally

reverberate at different levels, R1 and R2.

At this point, it appears that a network of neuronal sets designed in

this way exhibits three phenomena after a canonical string is presented.

These three phenomena do not occur together when an ill-formed stringis being processed. The phenomena are functionally and causally relatedwithin the given network architecture. Although the phenomena are func-

tionally related, they may be envisaged to be the basis of different cognitive

processes.

Phenomenon 1 is that the input units and the sequence detector connected

directly to the input units have ignited and therefore reverberate at a high

Page 202: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 202/331

10.5 Activity Dynamics of Sequence Detection 185

level. More generally, for each input unit activated by an element of the

input string, at least one group2 of sequence sets it is directly connected with

is in a high state of reverberation. If one or more sets of sequence detectorsconnected directly with an input unit are reverberating at high levels of R

(either R1 or R2)3, the input unit is called satisfied.Phenomenon 2 is that all reverberating input units the word webs act-

ivated by the string in the input are themselves at the highest level R1 of 

reverberation. Neuronal sets reverberating at R1 are called visible.

Phenomenon 3 is that all the sets that did ignite in response to parts of the

string in the input have finally participated in a continuous wave of ignitions

running through the network. This terminal wave of ignitions is initializedby the ignition caused by the last element in the input string. The ignition

wave is said to synchronize the neuronal sets involved.

Out of the three phenomena – satisfaction, visibility, and synchroniza-

tion – not a single one is met if the network in Figure 10.1 is presented with

the input BA. No synchronization took place as part of the processes caused

by the input. One of the input units is not visible: β is left at R2 rather thanR1. Also, both input units are not satisfied, because the only sequence set

they are connected to is primed rather than reverberating at a high level. Incontrast, all three conditions are finally met when the well-formed string AB

is processed. This simple example illustrates basic mechanisms of sequence

detection that are also relevant in the more realistic linguistic examples ex-

plained in the Excursuses.In summary, three criteria can be proposed for acceptance of an input

string by a network of neuronal sets:

(i) All input units activated by the string are satisfied

(ii) All of them are visible(iii) All neuronal sets activated are synchronized

If these criteria of satisfaction, visibility, and synchronization are reached,

the network is assumed to “accept” the string and then deactivate the entire

representation.

2 In this example, there is only one sequence set. However, in the next section, it is argued

that each input unit (representing a word or morpheme) must be connected to one or moresets of sequence detectors. These sets are supposed to be essential for disambiguating string

elements in the input.3 The motivation for allowing R2 in sequence detectors here becomes more obvious in fol-

lowing sections in which the processing of lexically ambiguous words is addressed. Briefly,

if a word is used twice and as member of different lexical categories, two sets of sequence

detectors are assumed to be reverberating, but only one of them is allowed to reverberate

at R1.

Page 203: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 203/331

186 Neuronal Grammar

10.6 Lexical Categories Represented in Neuronal Sets

10.6.1 Why Lexical Categories?

The primary task of a syntax network is to show a particular activity pattern

when input elements occur in a well-formed or grammatical string. The least

that the neuronal elements detecting a string AB must achieve is finding outthatPropositions(i)–(iii)arecorrect.(Recallthattheactivationofaneuronal

unit can be described as the verification of the statement that the entity the

neuronal element represents was present in the input of the neuronal unit

[cf. Section 6.1; McCulloch & Pitts, 1943; Kleene, 1956].)(i) A occurs

(ii) B occurs(iii) A occurs before B

Verification of Statements (i) and (ii) can be achieved by input units, the

neuronal sets that respond specifically if given morphemes, words or mean-

ingful word parts, are heard or read. As detailed earlier, a complex word,

such as switches, can be broken down into meaningful units, in this case switch and the suffix -es, that can realize the plural morpheme or the third

person singular present verb suffix. Therefore, input units are assumed to

respond specifically to either words or morphemes. If input units are con-

ceptualized as neuronal sets, it is clear how the knowledge that a given word

has occurred is kept in memory. The input unit α activated by a word ormorpheme A in the input verifies and stores the piece of information that A

occurred.

To make it possible for the network to decide (iii), whether A occurredbefore B, it appears straightforward to assume neuronal elements that spe-

cialize in sequence detection, such as the sequence detectors or sets proposed

earlier. The activity of a neuronal sequence detector would then mean veri-

fication of Statement (iii), that A occurred before B, by the network.One may therefore postulate sequence sets for word or morpheme se-

quences. Examples would be neuronal sets responding specifically to the

morpheme sequences “the machine” or “switch . . . on.” However, if each

possible word sequence or word pair sequence of a language were to berepresented by specific sequence sets, their number would be astronomi-cally high. An obvious solution to overcome this problem is suggested by all

grammar theories. The solution is to categorize words into lexical categories

and base serial-order algorithms on these grammatically defined word cate-

gories (see Chapter 7). The members of one of these categories can replace

each other in particular sentence contexts without changing the grammatical

Page 204: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 204/331

10.6 Lexical Categories Represented in Neuronal Sets 187

status of the sentence. If lexical elements, words, and morphemes, are

grouped into 100 or so lexical categories, the number of sequence detectors

necessary would be greatly reduced, because sequence sets could connectto a few category representations instead of an extremely large number of 

pairs of input units. This strategy may therefore offer a less costly solution

to the problem of sequence detection (see Chapter 9).Unfortunately, however, this solution implies that the basic task of decid-

ing whether an input sequence AB is congruent becomes more complicated

than indicated at the beginning of this section. It would now consist of ver-

ifying the following statements:

(i) A occurred(ii) B occurred

(iii) A belongs to lexical category a

(iv) B belongs to lexical category b(v) The element of class a occurred before the element belonging

to class b.

Statements (iii) and (iv) must still be spelled out in terms of neurons.

10.6.2 Lexical Ambiguity

The categorization of a word or morpheme into a lexical category may be

achieved by links connecting the neuronal set representing the lexical item

with neurons representing lexical categories. When a word is perceived,an input unit specializing in detecting this particular word ignites. It then

activates a strongly connected set that also receives strong connections from

sets representing other words of the same lexical category as the word thatoccurred in the input. In this case, categorization of the perceived word

would be straightforward.

However, the issue of assigning lexical categories to word forms is com-plicated by the fact that many words can be members of different lexical

categories. Here, it becomes necessary to distinguish between a word form

or morpheme at the surface and the surface element classified as a particular

type of lexical and syntactic unit. A word such asbeat can be categorized as a

nounorasaverb(the beat vs. to beat ). Words of this kind, that is, word formsthat can belong to different lexical categories, are called lexically ambiguous.A closer look shows that there are different homophonous verbs to beat that

differ in their valence, that is, the number of complements they require. The

issue of lexical ambiguity and that of homophonous verbs differing in their

valence is addressed in detail in Chapter 7 in the context of dependency

grammar.

Page 205: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 205/331

188 Neuronal Grammar

Following are more example sentences in which the word form beat  is

used as member of different lexical categories:

(1) The beat resonates.

(2) The heart beats.

(3) Betty beats Hans.

(4) Betty beats Hans with a stick.

(5) Betty beats Hans up.

In(1),beats isusedasanoun;in(2)–(5)beats isusedasaverb.In(2),onlyone

complement noun phrase is required, whereas another verbbeat , sharing the

word form with the item in (1) and (2), appears to require two complements,and in (4) even three. Finally, the verb beat  in (5) takes a verb particle as

its complement, in addition to two nouns. The different serial-order con-

straints of the homophonous words have their counterpart in different mean-ings. The word can be paraphrased by rhythm in (1), pulsates in (2), defeats

in (3), hits in (4), and hurts in (5).

The serial-order constraints of these different but homophonous words

can be characterized using syntactic rules, such as the dependency rules

(6)–(10).

(6) N (Det /*/)

(7) V1 (N1 /*/)

(8) V14 (N1 /*/ N4)

(9) V145 (N1 /*/ N4, Prep)

(10) V14p (N1 /*/ N4, Vpar)

Dependency rule (6) means that the word used as a noun requires one com-plement, a determiner to its left. Rule (7) says that the one-place verb re-

quires a noun in the nominative case to its left, and Rule (8) specifies an

accusative noun as an additional complement of a different verb type. In (9)and (10), two more verb types are defined with additional complements to

their right, a preposition or a verb particle. (For further explanation of the

abbreviations used and details about syntactic algorithms, see Chapter 7).

Evidently, the word form beat  is not the only one to which one of the

Rules (6)–(10) apply. Each of these rules operates on a large subcategory of nouns and verbs.

This consideration makes it obvious that assigning a word A to a lexical

category a is closely tied to specifying which kinds of words B1, B2, B3, . . . a r e

required to occur – and therefore usually occur – together with word A.

Extending this idea from words to morphemes, one can state that each

verb stem not only requires noun complements, but that it also requires a

Page 206: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 206/331

10.6 Lexical Categories Represented in Neuronal Sets 189

verb suffix Vs, which also depends on the case and number of the subject

noun. The verb subcategory definitions would therefore be as exemplified

in the formulas (11)–(14).(11) V1 (N1 /*/ Vs)

(12) V14 (N1 /*/ Vs, N4)

(13) V145 (N1 /*/ Vs, N4, Prep)

(14) V14p (N1 /*/ Vs, N4, Vp)

Verb complements such as affixes, nouns, and verb particles follow verbs,

although at different distances. Similar statements can be made for nouns

and other lexical categories. A nominative noun, N1, usually occurs after adeterminer, Det, and before the verb, whereas an accusative noun is usually

placed after an element of certain verb subcategories, as summarized by (15)

and (16).

(15) N1 (Det /*/ V)

(16) N4 (V, Det /*/)

There is, in principle, no limit for the level of detail one introduces in lex-

ical categorization. Distinctions can be made between different noun typesaccording to their grammatical case – nouns in the cases of nominative N1,genitive N2, dative N3, and accusative N4 – or ordinary nouns, N, from

proper names, Prop, and pronouns, Pro. More fine-grained lexical catego-

rization can also be motivated by the fact that the subcategories (e.g., nouns,

proper names, and prepositions) are subject to different serial-order restric-

tions. More fine-grained differentiation of lexical categories can also involvesemantic criteria, distinguishing, for example, between the objects that can

be the subject or object of beating. Clearly, more fine-grained distinctionsincrease the number of to be distinguished lexical categories. Current de-

pendency grammars propose approximately 100 different lexical categories

and subcategories.

The contrast between common nouns and proper names may be used as afinal example of a more fine-grained distinction of lexical categories. Articles

occur before common nouns but usually not before proper names, Prop. The

proper names used in the nominative and accusative cases may therefore be

characterized by (17) or (18). Standard nouns have been characterized by(15) or (16).

(17) Prop1 (/*/ V )

(18) Prop4 (V /*/)

Thisformulationisdifferentfromusualformulationsindependencygram-

mars. For formal reasons, grammar theories define the dependency relation

Page 207: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 207/331

190 Neuronal Grammar

as asymmetrical. If A is the dependent constituent and B is the “govern-

ing” constituent, the opposite is not possible. If the noun is the dependent

category and the verb governs it, the reverse is not the case. However, com-parison of the examples used here [e.g., (11) and (15)] shows that this is not

so in the present formulation. The dependency relation is assumed to bebased on reciprocal connections between word webs and sequence sets. For

this reason, it is defined as symmetrical in the present context. The noun and

the verb would therefore be considered to be interdependent, and the same

would apply for the verb and its suffix. One may want to speak about mutual

dependency here. If two lexical categories a and b are interdependent, the

dependency relation can be expressed by two rule types, (19) and (20).(19) a (/*/ b)

(20) b (a /*/)

The idea that lexical categories come in one canonical order is, of course,a simplifying assumption. For some of the serial-order regularities used here,

exceptions can easily be found (e.g., “Off went her hat”). Additional detail

mustbeintroducedtodescribethese,andthiswouldbepossiblebyspecifying

additional regularities by further algorithms.

10.6.3 Lexical Categories as Sets of Sequence Sets

Because lexical categorization of a particular word is so closely tied to the

question about which lexical categories occur regularly in the vicinity of the

element in question, it may be fruitful to answer the question of lexical cate-

gorization in terms of sequence regularities. For this purpose, a sequence set

can be conceptualized as a neuronal set specialized in detecting a sequencefeature. A sequence feature is a pair of elements of categories a and b in theorder “first a then b.” A sequence feature would be present in an input string

regardless of how long the lag is between the occurrence of the member of 

category a and that of the member of category b. The sequence detector

ignites only if a given sequence feature is present in an input string. After

its ignition, the sequence detector stays in a state of reverberation, thereby

storing the sequence feature in active memory.

A lexical category representation can now be defined as the union of several sequence detectors, as a set of sequence sets. Examples would be as

follows: The category of a nominative noun would be represented by two

sequence detectors, one detecting that the element in question followed an

article and the other examining whether it is followed by a verb. Accusative

nouns may be represented by the union of sequence detectors, one of which

Page 208: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 208/331

10.6 Lexical Categories Represented in Neuronal Sets 191

is sensitive to a transitive verb of a certain kind preceding the element in

question, and the other detecting a preceding determiner. The representa-

tion of the category of a transitive particle verb may include sets sensitiveto the word followed by a verb suffix, an accusative noun, and a particle.

Essentially, for each lexical category label occurring in a bracket of one of the postulated dependency rules, a corresponding sequence detector can be

postulated at the neuronal level.

Some of the postulated sequence sets connected to a particular word

representation would, so to speak, “look to the past” – that is, be sensitive

to particular inputs preceding the word, whereas others would “look for-

ward into the future” – that is, be sensitive to the word followed by otherelements. Thus, a lexical category a could be represented by both “future-

oriented” or “forward-looking” sequence features, a  f 1, a f 2, . . . , a f n, and

by “past-oriented” or “backward-looking” features, a p1, a p2, . . . , a pm. The

neuronal representation of the lexical category would include the corres-

ponding future- and past-related sequence sets, α f 1, α f 2, . . . , α f n, and α p1,α p2, . . . , α pm, respectively.

In the Rules (6)–(20), each lexical category label in the brackets would

be analogous to a sequence set. A lexical category label left to an asteriskbetween slashes corresponds to a backward-looking sequence set, and a

lexical category label to its right corresponds to a forward-looking sequence

set. It is clear from Rules (6)–(20) that words and lexical categories can

be characterized by varying numbers of sequence features, some categoriesonly exhibiting future-oriented features and others only showing regular

relationships with their past. This will be further illustrated in Section 10.6.5

in the context of Figures 10.3 and 10.4.

Distinguishing between past- and future-oriented sequence detectors im-plies a slight modification of the view put forward in Figure 10.1. There, only

one sequence detector was introduced for detecting the sequence AB. In thepresent proposal, this unit would be replaced by two sequence detectors, a

future-oriented detector α f  receiving strong input from the input unitα,and

a past-oriented detectorβ p strongly projecting to the word web β (Fig. 10.2).

In addition, strong connections fromα f  to β p and the corresponding weaker

backward connections must be postulated. Recall that all connections are

assumed to be bidirectional but asymmetrical. Essentially, this modificationamounts to a duplication of sequence detectors that does not change ba-

sic functional properties of the networks. The modification is introduced to

make descriptions of neuronal grammar easier.

A summarizing remark on the symbols used for referring to linguistic

and neuronal elements: Word forms and grammatical morphemes are

Page 209: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 209/331

192 Neuronal Grammar

labeled by capital Latin letters, and for lexical categories, lower-case let-

ters are used. Greek letters are used to refer to neuronal sets representing

words or morphemes. Sequence sets may be labeled by a lower-case Greekletter (indicating a lexical category the represented feature is part of) plus an

index. A “p” or “f” (indicating whether it is past- or future-oriented, respec-tively) and a number may appear in the index. Sequence detectors can have

several labels if they belong to different lexical category representations.

10.6.4 Neuronal Requirements of a Grammar Machine

The following paragraphs present an estimate of the neurobiological re-quirements of the envisaged machinery of lexical categorization and subcat-

egorization of words in a language.

In principle, the number of past- and future-oriented sequence detectorsper lexical category representation may be large, but it is, in fact, difficult

to find lexical categories with more than five complement categories. In de-

pendency grammars, verbs, which probably represent the lexical categories

with the largest number of complements, are sometimes assumed to have up

to four complements (Eisenberg, 1999; Heringer, 1996; Tesniere, 1959). Anexample would be the verb form “feed” in (21).

(21) Peter feeds Hans soup with a spoon

Counting the verb suffix “s” as an additional complement, the number of 

complements would be five in this case. Now, each input unit representing a

word form or morpheme can be strongly connected to all the sequence de-

tectors included in the representations of its possible lexical categories. Take

the extreme case of a five-complement verb, which is lexically ambiguousand can be classified, as a function of context, into one of five different lexicalcategories, each with another set of five complements. In this case, the input

unit of the word form would need connections to 25 sequence detectors. It

is not unrealistic to assume that for each word form representation, such a

small number of connections are established during language acquisition.

Assuming a language including 100,000 word forms or morphemes, each

of which is part, on average, of two out of 100 lexical categories, each defined

by five sequence features. To represent the vocabulary of this language, witheach word stored by a corresponding neuronal set, 100,000 input units would

be required one for each word. As mentioned, free pair-wise recombination

of the 100 lexical category labels results in 10,000 possible pairs. However,

the grammar network would not need most of these because most of the

possible sequences of categories would not occur regularly. To represent

Page 210: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 210/331

10.6 Lexical Categories Represented in Neuronal Sets 193

Table 10.2. Attempt at an elementary description of neuronal grammarat different levels. Linguistic entities are listed with their assumed

macroneuronal realization in neuronal sets. The number of neuronsinvolved at the microneuronal level is also estimated. According to thisestimate, the grammar circuits require surprisingly few neurons. Thefollowing assumptions underlie the computations: One neuronal setincludes 100,000 neurons. One language includes 100,000 words. Thewords of a language fall into 100 different lexical categories. Each lexicalcategory is characterized by 10 or less sequence features.

Linguistic Macroneuronal: sets Microneuronal: neurons

Word/morpheme 1 input unit each 105 neuronsSequence feature 1 sequencing unit 105 neurons

Lexical category <10 sequencing units <106 neurons

Lexicon 105 input units 1010 neurons

Grammar 103 sequencing units 108 neurons

107 connections between sets

the 100 lexical categories, each with a maximum of five complements, 500sequence detectors would be required. Doubling the number of sequence

detectors, as suggested for illustration purposes, would result in a maximum

of 1,000 sequence sets. Thus, 1,000 sequence detectors would be required

for representing a complete grammar in the way outlined here. The number

increases if more lexical categories are being postulated.

Nevertheless, the preliminary result is that the number of word or mor-

pheme representations would by far outnumber the number of sequence

detectors necessary. Assuming that the size of a neuronal set is 100,000 neu-rons, the maximum number of neurons of the vocabulary or lexicon would

be 1010 and the maximum number of grammar neurons making up sequence

detectors would be 109. These numbers of neurons are available for language

processing in the cortex. Table 10.2 summarizes the present estimate of the

“size” of the serial-order machinery. Note again the strikingly low number

of neurons required for wiring syntactic information.

The number of axonal connections between input and sequence sets

would probably be much larger than the number of neurons involved. Forconnecting the 100,000 input units to their proposed five category represen-

tations, each of which is made up of five sequence detectors, a few million

connections linking input and sequence sets would be required. Since most

cortical neurons have above 104 synapses, the large number of connections

should not constitute a problem. Note that not all neurons of a set need

Page 211: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 211/331

194 Neuronal Grammar

to connect to another set to establish a strong functional relationship be-

tween both sets. If there are 107 macro-level connections between neuronal

sets each constituted by 1,000 axons, the number of neuronal links wouldbe 1010, thus only a small percentage of the ∼1015 axons of the neurons

involved. Again, the proposal appears realistic.According to these estimates, a number in the range of 105 neuronal sets

is necessary for representing the vocabulary or lexicon of a language, whereas

the more narrowly defined grammar network, that is, the neuronal ensembles

specifically representing serial-order information, would be much smaller.

Only if a substantially higher number of lexical categories is postulated

(e.g., to 10,000 categories), the number of necessary sequence detectorswould reach the number of stored words or morphemes (105). Thus, an

increase in the estimate of the number of word categories and sequence de-

tectors leads only to a relatively small increase of the size of the required

network. The introduction of new sequence detectors responding to ad-

ditional features of a string opens up the possibility of representing more

word categories, including, for example, semantic subclasses of lexical cate-gories, in an economical way.

10.6.5 Lexical Disambiguation by Sequence Sets

A neuronal set representing a word form such as beat  or switch, which,

depending on within-sentence context, is part of one of several lexical cat-

egories, can now be assumed to be linked closely to all the neuronal setsrepresenting these alternative lexical categories. However, for each of its

occurrences, the word form should be classified as belonging to only one of 

the categories. This requires a mechanism deciding between the neuronalsets representing alternative lexical categories to which a given word can be

assigned. The mechanism can be realized by a regulation mechanism, for ex-

ample, mutual inhibition between the alternative representations of lexicalcategories. The most active lexical category representation would therefore

become fully active and the competitor category representations would be

suppressed.

More precise criteria that could allow for categorizing a word form A as

either a member of lexical category a or b could be the following: A could beclassified as a member of a if all sequence sets included in the representationof the a category have ignited a few time steps earlier, and if there is at

least one sequence detector included in the representation of b that did not

ignite. This allows only for deciding between sets of sequence detectors, each

of which includes specific sequence sets. However, if the representation of 

one lexical category representation includes the set of feature detectors of 

Page 212: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 212/331

10.6 Lexical Categories Represented in Neuronal Sets 195

Figure 10.3. Putative neuronal organization of lexical categories. An input unit representing

the word form switch is connected to three competing lexical category representations,as indicated by ovals in different tones of gray. The three alternative lexical categories arenominative noun, accusative noun, and transitive particle verb. Each category representationconsists of two or more sequence sets, indicated by small circles. Each of the sequencesets responds specifically to a different feature of the context, as indicated by questions.Inhibition between lexical category representations is provided by an inhibition mechanism.Modified from Pulverm ¨ uller, F. (2000). Syntactic circuits: how does the brain create serialorder in sentences? Brain and Language 71, 194–9.

the other [as would be the case for V1 and V14, cf. (7) and (8)] the decisioncriterion must be refined. For example, if the last inactive sequence detector

included in the representations of both a and b ignite at the same time, the

set with the larger number of sequence detectors wins.This or a similar “winner-takes-all” dynamics can be implemented by a

network of lexical category representations, each composed of one or more

sequence detectors if inhibitory connections are present between all repre-

sentations of lexical categories. Figure 10.3 sketches the putative neuronal

wiring of three homophonous words. Figure 10.4 presents the same wiringwith the more abstract labels used in Section 10.6.3. In this illustration, oneinput unit, the neuronal counterpart of the word form, is connected to three

sets of sequence sets. These sets are indicated by ovals in different shad-

ings of gray. The ovals are labeled by Greek letters, each of which would

be thought to represent one of the lexical categories the homophones can

be part of. Each set comprises two to four sequence detectors labeled by

Page 213: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 213/331

196 Neuronal Grammar

Figure 10.4. Putative neuronal organiza-tion of lexical categories: more abstractformulation. An input unit is connectedto three competing lexical category rep-resentations, α, β, and γ , each consist-ing of two to four sequence sets labeledby a Greek letter plus index. The setsof sequence sets organizing lexical cat-egories inhibit each other so that onlyone of them can become fully active to-gether with the input unit. (A more con-crete case is discussed in Fig. 10.4.)

Greek letters plus index. To guarantee inhibition between lexical category

representations, the sequence detectors would be connected reciprocally to

additional neuronal elements that would be connected among each other byinhibitory connections. This inhibitory wiring scheme is reminiscent of the

one proposed for corticostriatal circuits (see Fig. 5.2). The inhibitors would

be activated only if all the sequence sets included in their respective lex-

ical category representation were active at a time. Furthermore, this type

of between-category inhibition would become effective only if a word formactivates more than one lexical category representation at a time. If all se-

quence sets included in a lexical category representation reverberate at thesame time, they coactivate the inhibitor that, in turn, feeds excitation back

to the sequence sets. The input from the inhibitor would not be assumed to

be essential for maintaining enhanced activity levels in sequence detectors

[see principles (A3) and (A3’) below].

As emphasized, inhibitory interaction between alternative lexical cate-

gory representations of a word form is needed for disambiguation of lex-ically ambiguous word forms. Although no data are available that would

allow for narrowing down the possibilities for realizing mutual inhibition of 

lexical representations, a mechanism of this kind is a necessary component

of a neuronal grammar. This wiring scheme may have a possible brain basis

in corticostriatal connections and inhibition in striatum, but could also be

Page 214: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 214/331

10.6 Lexical Categories Represented in Neuronal Sets 197

realized by other direct or indirect inhibitory links between cortical repre-

sentations.

How would such a circuit actually solve the task of categorizing a lexicallyambiguous word form as a member of one of its possible lexical categories?

The “decision” should be made by a circuit including mutually exclusive setsof sequence detectors, each representing one of the possible lexical cate-

gories. To answer this question, consider again the more concrete example

of the word form switch. Depending on the context in which it occurs, it can

be classified, for example, as a noun in the nominative or accusative case, N1

or N4, or as a transitive particle verb, V14par. The characteristics of these

lexical categories can be expressed by dependency rules formulated earlierin this section, repeated here as (22)–(24) for convenience.

(22) N1 (Det /*/ V) Nominative noun

(23) N4 (V, Det /*/) Accusative noun

(24) V14par (N1 /*/ Vs, N4, Vpar) Transitive particle verb

This formulation in terms of syntactic algorithms can be translated into the

language of neuronal sets, as illustrated by Figure 10.3.The alternative lexical categories specified by (22)–(24) are represented

by sets of sequence sets, as indicated by ovals at different gray levels. Each

complement required by one of the three possible lexical categories of theword form has its corresponding sequence detector in the figure. The mutual

inhibition of lexical category representations ensures that only one of the

three alternative sets of sequence detectors is active with any occurrence of 

the word form switch.

One of the sequence feature detectors, the uppermost on the left in thefigure, responds to the word form switch that after an article (or determiner)occurred in the past. It ignites if a word form that can be used as an article

had occurred and the word form switch (or any other word form that can

function as a noun) follows it. On the upper right is a sequence set responding

to the word switch followed by a verb. This sequence detector can be thought

of as checking the near future for any word form that can be used as a verb.

Because it is being primed by the occurrence of  switch, it ignites when a verb

follows it. If both the past-oriented sequence set responsive to an articleand the future-oriented sequence detector targeting a verb have ignited, the

complete representation of the nominative noun category (oval in dark gray)

reverberates. The word form unit is therefore satisfied and assigned, by the

network, to the respective lexical category. This happens in response to a

string such as (25).

Page 215: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 215/331

198 Neuronal Grammar

(25) The switch broke.

The neuronal representations of the words in italics are assumed to be linked

tothewordwebof  switch. Similar processes take place if a verb and an articleboth precede the word form switch, as in sentence (26).

(26) Betty turned  the switch on.

In this case, two past-oriented sequence sets, primed by determiner and

verb, respectively, ignite when the word form switch occurs.

If the word form switch occurs after another word that can be used as a

noun, the sequence detector on the lower right ignites. Given the word formis also followed by a verb suffix, still another noun (in the accusative case)

and a particle, three more sequence sets sensitive to future events ignite

(after being primed by switch). These would be processes triggered by the

input string (27).

(27) The man switched the engine off.

In this case, the word form switch would ignite in close vicinity to the ignition

of all sequence detectors defining the neuronal representation of the lexicalcategory of a transitive particle verb.

In response to the example sentences, sequence detectors from nontarget

lexical category representations may also ignite. For example, when sentence

(27) is presented to the circuit fragment depicted in Figure 10.3, the sentence-

initial article primes the uppermost sequence detector on the left, whereasthe second word of the sentence, the noun man, primes the sequence set

at the lower left. The word from switch in the input ignites both primed

sequence detectors. Nevertheless, the further evidence accumulated whenthe additional words of the string (27) are processed activates all sequence

detectors of only one lexical category representation, that of the transitive

particle verb. Between-category inhibition is assumed to minimize activityin the competing lexical category representation.

It should be mentioned that this simple description glosses over a few facts

that may complicate the operation of a mechanism of this type. In the ear-

lier formulation in rules (22)–(24), different noun types marked for nomina-

tive or accusative case were distinguished. Making this distinction based onthe surface form of words is perhaps impossible in English, because nom-inative and accusative case are not overtly marked in this language. Fur-

thermore, the verb suffix may also not be realized; for example, in the first

person singular present case. In English, lexical subcategorization may pri-

marily rely on serial order information. Further relevant information about

lexical subcategories may be extracted from phonetic properties of spoken

Page 216: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 216/331

10.6 Lexical Categories Represented in Neuronal Sets 199

language. The distinction between grammatical cases is most obvious in

languages where these are overly marked, for example by specific inflectional

suffixes.After the input unit of a word has ignited and therefore reverberates, it

may be possible that a large number of future-oriented sequence sets, whichare part of representations of possible lexical categories of the word, are

primed. This activity is transmitted to all of the sets representing words of 

lexical categories that can be successors of this word. A rather large num-

ber of input units are therefore primed. Nevertheless, a successor word se-

lects and fully ignites only one of the sequence feature detectors connecting

its representation to that of switch

. If only one of the possible successorwords whose input units are primed occur in the input, the respective in-

put unit exhibits primed ignition and ignitions spread back to the input

unit of switch. In this case, the input unit is synchronized with (part of) the

verb representation. This happens until all sequence sets of the verb repre-

sentation have been ignited and therefore reverberate. Inhibition between

lexical category representation finally suppresses the competing categoryrepresentations.

These examples illustrate that in a network of neuronal sets, both for-ward and backward flow of activity can contribute to disambiguation and

classification of lexically ambiguous words into one of the lexical categories

they can be part of. The classification can be complicated; for example, if a

word exhibits multiple possible relationships with many other elements of astring, or if two word forms with the same lexicosyntactic properties occur

in the same sentence. Some of these problems are addressed in the more

formal reformulation of the proposal in the next section and in Chapters 11

and 12.The general idea put forward in this section is that the possible classi-

fications of a first element A in a string allow to “nominate” (by priming)potential successors B1, B2, . . . , which, if one of them is chosen (i.e., ig-

nites), in turn contributes to disambiguation of the initial element A, thus

“approving” one of its possible context-sensitive classifications. The idea is

very simple and is motivated by neuroscientific considerations. It appears

necessary and may be fruitful to spell out its possible brain basis in more

detail. The mutual nomination and approval of classifications of ambiguouselements occurring in a string requires reciprocal connections between their

context-sensitive representations.

NotethatthisideahassomerelationshiptowhatDilthey(1989),aphiloso-

pher, called thehermeneuticcircle–thatis,thedevelopmentofahigher-order

understanding of a symbol, or chain of symbols, on the basis of the processingof a second symbol whose interpretation is, in turn, influenced by the earlier

Page 217: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 217/331

200 Neuronal Grammar

symbols. In the present illustrations, only crude lexicosyntactic classifications

(noun or verb) and subclassifications (one- or two-place verb, nominative

or accusative noun) were made, but, of course, they have semantic implica-tions. The idea may well be extended to more fine-grained semantic distinc-

tions. Disambiguation of an earlier element is achieved by the occurrence of (a set of) later elements with which it has a regular serial-order relationship,

and conversely, the later element(s) can also be classified based on informa-

tion provided by the earlier element. For this to operate, activity caused by

symbols in the input must be assumed to reach at least the representation of 

possible successors of the symbol and must be allowed to flow back from the

successors’ representations to the representation of the original one. Thismechanism can be spelled out in the language of neuronal sets.

10.7 Summary: Principles of Neuronal Grammar

For making scientific progress, theorizing is necessary, and a newly developed

model needs to be illustrated in a concrete manner. To define functional

principles in a way that would make computer simulation and animations

possible, some of the assumptions must be made more precise. To this end,arbitrary decisions are sometimes necessary. To focus on a concrete example,there is no strong reason to assume, as proposed here, that activity of a

particular input unit only primes input units of its possible direct successors.

A two-step or higher-order priming process would also be possible, although

computationally more demanding. As a second example, it is necessary to

restrict activity in the network, but there are several options for choosing

the exact criterion for activating the threshold control mechanism.

Recall that it is the primary purpose of the present considerations todemonstrate syntactic competence in neuronal networks that may be housed

in the brain. (The fact that they could be organized slightly differently is

irrelevant for this purpose.) Consideration of alternative possible neuronal

mechanisms is, of course, highly relevant. It is possible, or even likely, that

details of the networks proposed here are incorrect as descriptions of actualcircuits in the brain. They can be modified once the necessary experiments

have been performed.

A few terminological remarks are necessary before the principles as-sumed to underlie the dynamics of the proposed circuits can be formulated.

The principles may reflect properties immanent to the human central ner-

vous system. Without making this assumption, one may prefer to considerthem axioms of the grammar networks proposed. Six axioms, (A1)–(A6), are

formulated, some of which will be reformulated later. The predicate S ( x, y)

specifies activity state x at time step y of neuronal set S. Although there are,

in principle, infinitely many different activity states (see Section 10.2), only

Page 218: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 218/331

10.7 Summary: Principles of Neuronal Grammar 201

four types are distinguished: rest (0), ignition ( I ), reverberation (R), and

priming (P ). As described in Section 10.2, different levels of reverberation

are called R1, R2, R3 (or R, R

, R

) and so on, with indices giving the rank inthe activity hierarchy. Larger numbers therefore mean lower activity levels.Priming levels are labeled in the same manner, as P 1, P 2, P 3 (or P , P , P ).

Now, a basic set of assumptions is as follows: The initial activity state of 

all sets is rest, 0. If there is no input to an inactive neuronal set, it stays in

state 0. If a set is reverberating, it stays at its reverberation state Ri – although,

as discussed in Section 10.2, its absolute level of activity may fall off with time.Ignition, I , and priming, P , are short-lived events. I is immediately followed

by reverberation (R1), and priming by inactivity (0). After an ignition, thereis a period of reduced excitability preventing too-frequent ignitions. The

double arrow (⇒) means “causes and is followed by.”

These assumptions can be expressed by (A1).

(A1) Spontaneous changes of activity in set S 

(i) S (0, t ) ⇒ S (0, t  + 1)

(ii) S (I , t ) ⇒ S (R 1, t  + 1)

(iii) S (P i , t ) ⇒ S (0, t  + 1)(iv) S (R i , t ) ⇒ S (R i , t  + 1)

(v) Refractoriness: If S (I , t ), then S cannot ignite at t  + 1 or at t  + 2.4

The effect of priming at time t has vanished one time step later, at t + 1.

Therefore, input at t  to a set exhibiting state P  but receiving no additional

input at t essentially has the same effect at t  + 1 as input to a set at rest.Spontaneous activity changes as summarized in (A1) are one factor deter-

mining network dynamics. There are four possible causes of activity statesin neuronal sets that will be considered.

(a) External input to input units.

(b) Activity flow between sequence sets and input units.

(c) Interaction between neuronal sets and threshold regulation.

(d) Inhibition between sets in the case of processing of ambiguous words.

Input to the network from outside by an element of an input string is

called external stimulation, E. External stimulation of an input unit causesit to ignite and reverberate later. The primed ignition I of an already highly

active unit contrasts with the stronge full ignition I ∧ of a previously inactive

unit. S (E, y) denotes the input E at time y to set S.

4 This principle overrides the principles formulated by (A3) and (A4):Immediately after an

ignition, a set cannot go back to I .

Page 219: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 219/331

202 Neuronal Grammar

(A2) Activity change in an input unit S  caused by external stimulation

(i) S (0, t ), S (E , t ) ⇒ S (I ∧, t  + 1)

(ii) S (P 1, t ), S (E , t ) ⇒ S (I , t  + 1)(iii) S (P i , t ), S (E , t ) ⇒ S (I ∧, t  + 1)

if i > 1

(iv) S (R 1, t ), S (E , t ) ⇒ S (I , t  + 1)

(v) S (R i , t ), S (E , t ) ⇒ S (I ∧, t  + 1)

if i > 1

The difference between full ignition I ∧ – as specified by (A2) (i), (iii), and

(v) – which activates threshold regulation [see (A6)] and primed ignition I [cf. (ii) and (iv)] which has no additional global effect on network dynamics,is important.

Clearly, if priming or reverberation levels are very low, their effect be-

comes negligible. Therefore, the principles addressing priming and reverber-

ation effects on ignition must be specified with regard to the activity levels

for which they apply. Regarding principle (A2), this cut-off is assumed to be

below P 1 and R1 – that is, only the highest levels of priming or reverberation

contribute to neuronal processing in the network as specified by (A2). Also,slightly lower levels of activity (P 2 and R2) are assumed to play a role in the

interaction of neuronal elements as specified by (A3) and (A4). Levels of 

priming and reverberation below R2 and P 2 (and sets exhibiting only such

low levels) are called invisible.Both input and sequence sets, receive internal stimulation – that is, ex-

citation through connections from other sets. A long arrow (→→) is used

to indicate a bidirectional link with stronger connections in the direction of 

the arrowhead. Recall that direction and strength of connections betweeninput and sequence sets carry information about the input strings that are

accepted by the network.Ignition of a set spreads through all connections to sets already exhibiting

a high activity level (i.e., reverberation or priming at rank 1 or 2). If Sq is less

active or at rest when set S p ignites or reverberates, activity of S p primes Sqonly if the connection from S p to Sq is strong. In this case, priming affects not

only strongly and directly connected sequence sets, but priming is further

communicated to the next input unit also.

(A3) Activity changes caused in Sq through a connection between Sp and Sq

(i) S p (I , t ), S q (R i , t ) ⇒ S q (I , t  + 1)

only if i = 1 or i = 2

(ii) S p (I , t ), S q (0, t ) ⇒ S q (P 1, t  + 1)

only if S p→→S q

Page 220: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 220/331

10.7 Summary: Principles of Neuronal Grammar 203

(iii) S p (R i , t ), S q (0, t ) ⇒ S q (P i , t  + 1)

only if S p→→S q(iv) S 

p(P 

i , t ), S 

q(0, t ) ⇒ S 

q(P 

i , t  + 1)

only if S p→→S q and S p is a sequence set

Spreading of activity through connected sets needs time. If external stim-

ulation reaches an input unit, it ignites one time step later, and during thefollowing time steps, priming is exerted on its future-oriented sequence set,

which in turn prime connected past-oriented units projecting onto further

input units. Thus, there is a lag of four time steps between the occurrence

of a symbol in the input and priming of the representations of its possiblesuccessors in the string. Only if three or more time steps separate the oc-

currence of two consecutive symbols in the input is the priming machineryeffective.

To pronounce a word consisting of one syllable, it takes approximately

150 ms or more although some inflectional affixes may be realized as even

shorter events. To allow for three time steps between the activity waves

caused by two one-syllabic words that directly follow each other, the “time

steps” assumed for the present syntactic derivations need to correspond to50 ms or less of real time, corresponding to a frequency of 20 Hz or higher.

This applies, for example, to the spreading of priming from one input unit to

its sequence set and to additional input units, or to the spreading of ignitions

from one set to a neighboring primed set. There are also longer-lasting ac-

tivity processes in the present neuronal grammar machinery. For example,

the backward spreading wave of ignitions representing an entire sentencewould be a much longer-lasting process, and importantly, the reverberation

of a set is assumed to last at least for tens of seconds.In the derivations shown next, it takes up to seven time steps after activa-

tion of an input unit by external stimulation until all relevant computations

have been performed. Only then will the system have reached a steady-state

in which spontaneous activity changes have come to a rest. To avoid effectsthat are related to the exact timing of incoming words – pauses or slight

variation in the speed of speaking not considered here – the next input from

outside should reach the network when it has settled into a stable state, that

is, only after seven time steps. Thus, to be on the safe side, it may be bestto conceptualize the time steps in the present computations as even shorter

intervals, of some 20–30 ms of real time in real networks. This correspondsto a frequency of ∼30–50 Hz.

One remark relevant to subsequent example simulations is necessary

here: If there are several past-oriented sequence sets in a lexical category

representation, then (A3iii) and (A3iv) must be modified as follows: Priming

Page 221: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 221/331

204 Neuronal Grammar

of an input unit of a possible member of such a lexical category is achieved

only if all the past-oriented sequence sets of the lexical category representa-

tions are primed. The level of priming is calculated as the average of activitylevels exhibited by the sets exerting priming on the input unit. If the average

is not a natural number, the next smaller number is chosen.More complicated issues in the activation dynamics of the proposed syn-

tactic circuits must be addressed. Imagine that two sets feed into a third

one. How would their activity states interact with the recipient to yield new

states?

If a given set Sr  is inactive but receives excitatory input from two other

sets S p and Sq, simultaneous input through both connections may summate.To change the state of a sequence set from inactivity to reverberation, at

least two simultaneous inputs are necessary. One input caused by an igni-

tion is necessary and the other input may also be ignition related, or may be

mediated by reverberation or priming communicated through strong con-

nections.

(A4) Activity changes caused in S r  through connections with S p and S q

(i) S p (I , t ), S q (I , t ), S r  (0, t ) ⇒ S r  (I , t  + 1)(ii) S p (I , t ), S q (R i , t ), S r  (0, t ) ⇒ S r  (I , t  + 1)

only if i = 1 or i = 2, and S q→→S r (iii) S p (I , t ), S q (P i , t ), S r (0, t ) ⇒ S r  (I , t  + 1)

only if  i = 1 or i = 2, and S q→→S r , and S qis a sequence set

Thus, two simultaneous excitatory inputs, one of which must be an ignit-

ion, causes a given set to be “switched on” – that is, to ignite and reverberatelater. In this case, R and P again exert their influence only through strong

connections, and only if their activity rank is high (R1 or R2; P 1 or P 2).

Only if ignitions in two sets are exactly synchronous can they cause an

additional ignition in a connected third set according to principle (A4i). Be-

cause the duration of words in spoken sentences and the delays betweenthem (and the scanning times for individual words during reading) can vary

substantially without affecting sentence comprehension, it would not be wise

to base the network’s grammaticality judgments on variations of timingof neuronal activation processes caused by the speed of the language in-

put. Therefore, only axioms (A4ii) and (A4iii) are relevant in the following

derivations. As mentioned, short time steps are assumed for the computa-tions in the network so that the network settles into a stable state by the time

a new input unit is activated.

Global network dynamics may exert an additional important influence

on the computations performed by syntactic circuits. Threshold regulation

Page 222: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 222/331

10.7 Summary: Principles of Neuronal Grammar 205

is activated by full ignition [see axiom (A2i)] and leads to reduction of all

levels of reverberation from Ri to Ri+1. Additional activations of threshold

regulation reduce the level toRi+

2,Ri+

3, and so on. Levels of priming areaffected in the same way. The regulation mechanism may also enhance neu-ronal activity if its level is low. In the present formulation, reverberation and

priming levels are just relabeled so that the highest levels are always R1 or

P 1. Nevertheless, lower activity levels are not adjusted so that not all levels

between the maximal and minimal level are necessarily present at all times

(e.g., the only states may be R1 and R3).

(A5) Threshold regulation

(i) if a full ignition I ∧ of an input unit happens at t , then for all reverber-

ating S :

S  (R i , t ) ⇒ S  (R i +1, t  + 1)

(ii) if there is no ignition or reverberation at R 1 at t, then - for all rever-

berating S:

S  (R i , t ) ⇒ S  (R i −1, t  + 1)

Because priming is caused by reverberation [as specified by (A3iii)] and by

ignition (A5), implies parallel adjustment of priming levels.As detailed in Section 10.6 (the section on lexical representations), alter-

native lexical categories of a given word or morpheme must exclude each

other. The postulated inhibition mechanismis assumed to exclude simultane-

ous ignition of two lexical category representations connected with one inputunit. Also, if part of a lexical category representation ignites or reverberates

at R1, activity of competing lexical category representations connected with

the same input unit are assumed to be reduced. The respective input unit

must be active (at I or R1) and thereby enable category inhibition.

(A6) Inhibition between two lexical category representations α and β con-

nected to an active input unit S  that ignites or reverberates at R 1:

(i) Ignition of S  can only spread to sets included in either α or β. The

most strongly activated representation wins.

(ii) If one or more sets included in α but not in β ignite at t , then no set

included in β and not included in α can exhibit I , P 1, or R 1 at t  + 1.

(A6i) is formulated without giving an exact criterion for calculating activ-

ity levels of lexical category representations, because it is only occasionallyrelevant in the circuits here (cf. Section 10.5.4). To decide which already ac-

tive alternative category representation, α or β, ignites, the average activity

level of the active sets in α and β can be chosen as an additional criterion.

If activity strength is calculated by averaging, three sets at R1, P 1, and R2

have a higher average level (1.3) than two sets at R1 and R2 (average: 1.5).

Page 223: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 223/331

206 Neuronal Grammar

If both competing category representations exhibit the same average level of 

preactivity, ignition spreads to the larger set (the one with the larger number

of sequence sets).In Chapter 12, the principles formulated by (A2)–(A6) are modified to

improve the present version of a neuronal grammar.

It is important that principles (A1)–(A6) are thought to reflect biologicalproperties of the nervous system described at the level of neuronal ensembles

and their interactions. Some of them are closely related to the established

neuroscientific principles discussed in Chapter 2 or to the mechanisms out-

lined in Chapters 5, whereas others are based primarily on brain-theoretical

postulates (see earlier discussion). Principles (A1)–(A6) can be viewed ashigh-level functional consequences of neurophysiological properties of neu-

rons and the way they are connected neuroanatomically in the human brain.

If this view is correct, the principles reflect neuronal structure and function

as determined by the human genome.

Page 224: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 224/331

CHAPTER ELEVEN

Neuronal Grammar and Algorithms

This chapter addresses the question of how to translate grammatical algo-

rithms into the language of neuronal sets.

11.1 Regular Associations, Associative Rules

There has been some discussion about the question of whether the humanmind and brain use neuronal principles and connections for processing gram-

matically related information, or whether it uses rules and algorithms spec-

ified by grammar theories (Elman, Bates, Johnson, Karmiloff-Smith, Parisi,

& Plunkett, 1996; Pinker, 1994). This proposal suggests that these positions,

which are sometimes considered to exclude each other, are, in fact, bothcorrect. This is not meant in the sense that there are two modules or sys-

tems, one for neural networks and the other one for rule algorithms (Pinker,

1997), but in the sense that rules are abstract descriptions of the neuronalmachinery, as they are, without any doubt, descriptions of aspects of human

behavior and action (Baker & Hacker, 1984).

If rules and algorithms are adequate descriptions of aspects of human be-havior and action, they must have a basis in neuronal structure and function.

As stressed in the discussion of the McCulloch–Pitts theory (Section 6.1),

a neuronal network can be reformulated using calculus or by a logical

formula. It is therefore reasonable to ask which putative neurobiological

counterparts exist for syntactic rules and, conversely, how a neuron circuitsensitive to serial order can be adequately described algorithmically.

The cortex is an associative memory and would therefore be difficult to

imagine it ignoring the correlation of words and morphemes in its input.

Rather, it may use the information about correlation of word pairs for set-

ting up sequence sets and their connections to the representations of word

groups whose members can be replaced by each other in a given context.

207

Page 225: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 225/331

208 Neuronal Grammar and Algorithms

The neuronal grammar wired up in this manner may finally exhibit prop-

erties very similar to aspects of rule descriptions in grammar theories. One

type of grammar model, dependency syntax, can, after some modification,be translated straightforwardly into the language of neurons, as has earlier

been suggested and is elaborated further in this chapter. Neuronal grammarmay be rooted in and limited by general principles immanent in neuronal

function.

From one viewpoint, each individual neuron performs a logical compu-

tation (Section 6.1), and therefore greater neuron ensembles would do so

as well. Strongly connected neuron ensembles are particularly well suited

to represent the elements of logical or algorithmic operations, because theyexhibit discrete activity states. The functional web or neuronal set either

ignitesordoesnot.Thisfunctionaldiscretenesscanbeassumedtobepresent,

although in different contexts of background activity, the exact boundaries

of the neuron set may slightly vary.

The equivalence of algorithmic descriptions and neuronal circuits can be

illustrated by spelling out linguistic rules and regularities in the languageof artificial neurons. For example, the rules of regular past-tense formation

and the regularities underlying past-tense formation of irregular verbs canbe contrasted using a neuronal automaton, and individual connections and

connection bundles can be correlated with the storage of information about

rules and exceptions (see Chapter 6; Pulvermuller, 1998). This chapter fur-

ther illustrates the equivalence of algorithmic descriptions and networks byproviding fragments of neuronal grammar depicted as neuronal circuits, and

also as algorithms.

The close relationship between a description in terms of syntactic algo-

rithms and a description of properties of syntactic circuits is also evidentfrom the summary of Chapter 10. There, principles underlying dynamics

of neuronal grammar were presented in the form of axioms (A1)–(A6). Areader who prefers algorithms to neuronal networks is free to take these ax-

ioms to express assumptions simply that allow for economical modeling of 

abstract syntactic processes. From a neurobiological perspective, however,

these same axioms are, in the best case, correct descriptions of principles

of neuronal dynamics in the human brain. In this case, the algorithmic and

neuronal formulations would be near equivalent.Axioms (A1)–(A6) may formulate what all grammar circuits have in com-

mon: their putative common ground in properties of the neurobiological

machinery. In this chapter, the focus is on the specific adjustments necessary

for representing a particular language. To speak a language, one must not

only have a largely intact brain, but must one also have learned words, their

assignment to lexical categories and subcategories, and the restrictions that

Page 226: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 226/331

11.2 A Formalism for Grammar Networks 209

apply to ordering them in time. Clearly, axioms (A1)–(A6) have implications

for the possible temporal ordering, but they also allow for some language-

specific variation. The assumption is that the latter can be learned on thebasis of Hebbian associative learning principles.

11.2 A Formalism for Grammar Networks

As already emphasized, syntactic circuits in the human brain, the neuronalside of grammar, can be characterized by

   r

neuronal and grammatical principles (see Section 10.)   r formulas representing language-specific regularities and wiring.

This section shows how the formulation can be adjusted so that for each

formal statement of the second type, there is a neuronal circuit or con-

nection pattern to which it refers. The assumption is that there are 1 : 1

correspondences between linguistic representations and neuronal entities(e.g., neuronal sets) and between linguistic processes and neuronal processes

(Pulvermuller, 1992).As mentioned, the formulas representing linguistic regularities (and

wiring schemes) use symbols for terminal elements, which refer to word

forms and affixes actually occurring in sentences (cf. Chapter 7). In addi-

tion, nonterminal symbols for lexical categories (referring to word and affixtypes) are used. As detailed, words and morphemes, the terminal elements,

have their putative organic basis in word webs or input units, whereas lexical

categories, one type of nonterminal symbol, are realized as sets of sequence

sets – “sequence supersets,” so to speak.What follows now is one possible statement of neuronal grammar in terms

of algorithms. Three types of formulas are specified: assignment, valence,

and sequence formulas.

Assignment formulas specify the correspondence between word forms

or affixes and lexical categories. This many-many assignment function (Bar-

Hillel, Perles, & Shamir, 1961) is specified by a set of assignment formulas,

each giving one or more lexical elements and the lexical category to which

they belong. A double arrow (↔) is used to denote the assignment relation.

(1) A, B, C, . . . ↔ a

By (1), the lexical elements A, B, C, and so on are assigned to lexical category

a. Each lexical element, word, or morpheme can be assigned to more than

one lexical category.

Page 227: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 227/331

210 Neuronal Grammar and Algorithms

Assignment formulas are proposed to be organized in the brain by recip-

rocal but asymmetrical connections between input units and a collection of 

sequence sets.Valence formulas, included in a language-specific neuronal grammar,

specify properties of a lexical category in terms of sequence features.

(2) a (p1, p2, . . . , pm /∗/f 1, f 2, . . . , f n)

By (2), lexical category a is characterized as having m backward sequence

features (briefly, backward features) specifying categories regularly occur-

ring within a sentence before an element of category a occurs. The n forward

sequence features (briefly, forward features) specify what must be expectedin the future given an element in the input is classified as member of lexicalcategory a. The asterisk between slashes indicates the relative position of 

the element assigned to a. This is similar to giving the valence of a lexical

category and listing how many of the complements must occur to the left and

right of a member of the category. Therefore, (2) is called a valence formula.

Note that no order of the preceding elements pi is implied, as there is no

preferred order of the elements f  j  following the member of the category a.The neuronal substrate described by valence formulas would be the

backward- and forward-oriented sequence sets connected to input units of a

certain kind. These input units frequently share their neighbors in sequences

of ignitions. They represent members from a given lexical category.

Sequence formulas specify individual sequence features and have thefollowing form:

(3) a ( f i )→→ b (p j )

The long arrow (→→) can be read “connects to.” In neuronal grammar,

this type of connection provides an important mechanism of serial ordering.

Sequence formulas have their putative neuronal equivalent in reciprocal but

asymmetric connections between sequence sets.The following two sequence formulas would be relevant for storing the

knowledge that certain types of English nouns and verbs follow each other

[see (4)], and that certain types of English verbs are followed by a particle

[see (5)].

(4) N (f )→→  V (p)

(5) V (f ) →→  Vpart (p)

Sequence formulas can be inserted into valence formulas to yield a de-

scription of lexical categories in terms of which other categories are required

Page 228: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 228/331

11.3 Some Differences Between Abstract and Neuronal Grammar 211

to the left and right. For example, inserting formulas (4) and (5) into (6) yields

formula (7).

(6) V (p /∗/ f )

(7) V (N /∗/ Vpart)

Formula (7) implies that an element categorized as an English verb (V)

of a particular subtype can come with a preceding noun (N) and another

element following it that can be categorized as a verb particle (Vpart).

Clearly, theverb category meant here is an intransitiveparticle verb (V1part).

Derivations of sequence and valence formulas such as (7) are called rules

of neuronal grammar and resemble valence rules used in dependency gram-mars (see Chapter 10 for discussion of common features and differences).

Assignment formulas can also be inserted into valence formulas or into

rules. By inserting the assignment formula (8) into Rule (7), the word-specific

Rule (9) can be obtained.

(8) get ↔ V

(9) get (N /∗/ Vpart)

Rules of neuronal grammar provide descriptions of sentence structures.

Such structural descriptions are obtained by inserting valence formulas or

rules into each other. The following set of abstract rules provides an example:

(10) c (b /∗/ e)

(11) b (a /∗/)

(12) e (d /∗/)

Inserting (11) and (12) into (10) yields bracketed complex rule (13).

(13) c [b (a /∗/) /∗/ e (d /∗/)]

The implications of (13) are the same as those of (11) and (12) applied after(10).

11.3 Some Differences Between Abstract

and Neuronal Grammar

These examples of insertions of formulas into formulas may remind one of 

what is possible with standard grammar algorithms. One may therefore want

to state that conventional grammars, for example, context-free rewriting

systems or dependency grammars (see Chapter 7) can be reformulated by

Page 229: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 229/331

212 Neuronal Grammar and Algorithms

a language-specific version of neuronal grammar. However, this would be

incorrect for the following reasons.

First, a rewriting system allows for multiple use of the same rule and lex-ical category representation. A rule can be used recursively in conventional

grammar models. This is not so in the present neuronal framework, in whicha given lexical category representation can be used only once in the course

of the analysis of a sentence. Chapter 12 addresses this issue and refines the

notion of a neuronal grammar so that repeated elements can be processed.

Furthermore, to provide evidence that neuronal grammars can compete with

traditional grammars in the description of natural languages, it appears of 

particular interest to investigate whether the former are capable of process-ing string types that are outside the reach of finite-state grammars, but are of 

particular relevance in the description of natural languages, namely strings

with center embeddings. This problem is taken as a test case for an extended

version of neuronal grammar in Chapter 12.

The second reason is that it is always possible to introduce a rule such as

(14) into an algorithm, such as the rule defined by (10)–(12).

(14) b (/∗

/ e)

This implies that, in a chain abcde defined by (13), there would be an addi-

tional relationship between the second and the last elements (b and e). On

the background of grammar (10)–(12), this would imply that an additional

serial-order constraint is introduced. The number of strings the grammar

accepts would become smaller.

There is no easy way to insert (14) into (13). Certainly, the new rule can-

not be inserted in the same manner as (11) and (12) are inserted into (10).This is because all elements defined by (13) are already in the formula, and

what would need to be added by (14) is an additional relationship between

temporally and spatially distant elements. If the rule derivation is repre-

sented in a two-dimensional graph (or tree; see Chapter 7), the addition of 

(14) to the system leads to crossings of lines in the graph, thus violating abasic principle of many grammar theories. In neuronal grammar, the multi-

dimensional relationships between elements of a syntactic string can be

captured, although it is not easily captured by bracketed formulas ortrees.

The emphasis is on that the circuit side has the same feature. Also, the

circuit side of neuronal grammar is not subject to the restriction of two-dimensional graphs. The circuits are diagrams realizing all the entities speci-

fied in the formulas of a fragment of neuronal grammar. Each lexical

element is represented as an input unit, each lexical category as a set of 

sequence sets, and each directed link defined by sequence formulas is

Page 230: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 230/331

11.3 Some Differences Between Abstract and Neuronal Grammar 213

realized by an asymmetrical reciprocal connection between sequence sets. In

principle, the lines symbolizing neuronal links may cross, although one may

believe that it is preferable to choose displays in which unnecessary crossingsare avoided.

No general statements are made here about the generative capacity of neuronal grammars except the following: Rule (15) specifies the set of 

strings (16).

(15) b (a /∗/ c, d)

(16) {abcd, abdc}

The addition of a second rule, (17), increases the size of the set of possible

strings quite considerably to the set listed under (18).

(17) a (/∗/ f)

(18) {afbcd, abfcd, abcfd, abcdf, afbdc, abfdc, abdfc, abdcf }

Instead of two string types eight would now be included. This is because

only the slashes (not the commas) in valence formulas and rules indicatesequential order. Note that this is a major difference between dependency

systems and neuronal grammars of the present formulation. In dependency

systems, a rule of the form “b (a, ∗ , c, d)” implies sequential order of allfour constituents so that only one string form – namely, abcd – would be

generated or accepted.

To exclude some of the eight string types defined by Rules (15) and (17),additional rules would be necessary; for example, (19), which again restricts

the set of possible string types.

(19) f (/∗/ c)

(20) {afbcd, abfcd, afbdc, abfdc}

The ability of a set of neuronal grammar rules to produce different sequences

of the same lexical categories could be helpful for modeling properties of 

languages. There are sentence types in which the serial order of certain con-

stituents is not constrained. Examples in English include double object con-

structions and sentences with transitive verbs that require particles, wherethe ordering of the two complements can be changed, as shown by example

sentences (21)–(25).

(21) Betty puts a hat on.(22) Betty puts on a hat.

(23) She rides on a bus to Bergamo.

(24) She rides to Bergamo on a bus.

In both cases, the sentences could be modeled by a rule such as (15) [repeated

here as (25) for convenience].

(25) b (a /∗/ c, d)

Page 231: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 231/331

214 Neuronal Grammar and Algorithms

Filling this abstract example with the concrete lexical categories used in

(21)–(24) results in (26) and (27), respectively.

(26) V14p (N1 /∗ / N4, Vpart)(27) V1PP (N1 /∗ / Prep1, Prep2)

In both cases, the verb requires two elements in its near future – accusative

noun plus particle in one case and two different prepositions (Prep1, Prep2)

in the other. The order in which these must follow is, however, not fixed

in language use. To allow this same freedom in the formalism, valence for-mulas with free placing of the elements to the right of the asterisk can be

advantageous.As emphasized earlier, this formulation has some similarity to traditional

syntactic theories, namely to dependency grammars and valence theory

(Gaifman, 1965; Hays, 1964; Heringer, 1996; Tesniere, 1953). The underlying

idea that members of lexical categories require complements (and can actu-ally be defined as requiring complements) that are members of other lexical

categories was expressed and systematically treated in Tesniere’s work, but

related ideas have been crucial in other grammar theories as well (e.g., cat-

egorical grammar; Ajukiewicz, 1936; Lambek, 1958, 1959). Formulas repre-senting connections between input and sequence sets (assignment formulas)are analogous to assignment rules or lexicon rules defining the lexical cate-

gories of words and morphemes. Rules of this kind are probably specified in

all major grammar theories. Descriptions of lexical category representations

(valence formulas) are inspired by and are somewhat similar to statements

specifying the valence of lexical categories and their dependencies, which are

typical of dependency systems (see Chapter 7). There is no obvious analo-

gon in grammar theories to the formulas representing connections betweenneuronal sets (sequence formulas).

11.4 Summary

In conclusion, the neurobiological grammar circuits proposed here have an

equivalent formal description. This formalism is not equivalent to standard

rewriting syntax or dependency theory, although it exhibits family resem-

blance with both. The main difference is that each formula refers to a neu-ronal device. Any attempt to simply reformulate the class of neuronal gram-

mars in the languages of phrase structure or dependency grammars results

in algorithms that produce trees with projection lines that cross or graphs

with loops, something prohibited in these grammar theories. It may be that

the present neurobiologically motivated proposal of syntactic algorithms can

open new perspectives for grammar theories.

Page 232: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 232/331

EXCURSUS TWO

Basic Bits of Neuronal Grammar

How does neuronal grammar operate? The following examples further il-

lustrate the activation processes taking place in a network of neuronal sets

during perception of congruent or grammatically well-formed and incon-

gruent or ill-formed strings. This excursus aims to illustrate the principleddifference in network dynamics between the processing of congruent and

incongruent word strings, and further aims to introduce illustration schemesfor network dynamics that are used in later sections of the book (see E3–E5

Chapters 11, 13).

Although the general mechanism of serial-order detection, mediated se-

quence detection by sequence sets, is simple, the interaction of several neu-ronal sets can become quite complex. To make their activity dynamics easy

to overlook, two strategies are used to illustrate processes in grammar net-

works. One strategy is to list activity states of all sets contributing to the

processing of a string at each point in time when a string element is presentin the input and shortly thereafter. Activity dynamics are therefore pre-

sented in the form of tables. A second strategy is to present the simulationsas animations. The animations, including illustrations of the three examples

presented in this excursus, are available on the Internet at this book’s ac-

companying web-page (http://www.cambridge.org).

E2.1 Examples, Algorithms, and NetworksStrings such as (1), (2), or (3) could be taken as examples for illustrating the

function of a simple grammar network.

(1) Betty get up.

(2) Betty switches . . . on.

(3) The set ignites.

215

Page 233: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 233/331

216 Basic Bits of Neuronal Grammar

These strings are grammatical and one may envisage a grammar network

to “accept” them. This means that the presentation of the word chains

causes activity dynamics in the network that satisfies certain criteria. Cri-teria for string acceptance in neuronal grammar networks are discussed in

Chapter 10 and are dubbed satisfaction, visibility, and synchrony. In this ex-cursus, simple illustrations of network dynamics present in a neuronal gram-

mar during and after presentation of strings of these types are presented and

discussed.

The type of grammar network considered here accepts strings such as

(1)–(3) because it includes two sequence detectors, one of which responds

to the first and second elements of the respective string while the otherresponds to the second word followed by the third. It is clear that suchsequence detectors would not only respond to these particular words, but

also to sequences of elements of larger word groups (see Chapters 9–11).

Although the first word pairs of sentences (1) and (2) could be detected by

the same sequence set, different sequence sets would probably be involved

in the processing of all other word-pair sequences.If the order of these words is being changed, one may feel that the re-

sulting strings are less common than the strings (1)–(3). This is illustrated bystrings (4)–(9).

(4) Up get Betty.

(5) On . . . switches Betty.

(6) Ignites set the.

(7) Up Betty get.

(8) On . . . Betty switches.

(9) Ignites the set.

Not all of strings (4)–(9) are incorrect or grammatically deviant. However,

it is assumed that there is one type of grammar machine that accepts strings

such as (1)–(3) and does not respond in the same manner to (4)–(9). Again,

this is for exposition purposes. It is possible to introduce additional networkelements whose interaction may yield acceptance of other combinations of 

these words as well. For example, strings (5) and (8) may be accepted because

of such to be added sequence detectors.

A network composed of neuronal sets that accepts the well-formedstring (1) but rejects alternative combinations of the same words is shown in

Figure E2.1. The grammar depicted in Figure E2.1 can be rewritten by thefollowing set of formulas:

(10) Betty ↔ N (noun)

(11) get ↔  V (verb)

Page 234: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 234/331

E2.2 Grammar Circuits at Work 217

Figure E2.1. Elementary grammar net-

work. Three word webs are connectedby two sequence sets. All simulations inExcursus E2 are based on this circuit.

(12) up ↔ Vpart (verb particle)

(13) N (/∗/ f)

(14) V (p /∗/ f)

(15) Vp (p /∗/)

(16) N (f)→→  V (p)

(17) V (f) →→  Vpart (p)

There is correspondence between parts of the circuit and formulas

(10)–(17). The relevant input units are defined with regard to their respec-

tive category labels by assignment formulas (10)–(12). Valence formulas

(13)–(15) list the connection of each input unit to backward- and forward-

looking sequencing units, and sequence formulas (16) and (17) have theirnetwork counterpart in the two horizontal arrows connecting two adjacentpairs of sequence sets.

E2.2 Grammar Circuits at Work

Table E2.1 presents a sketch of activity dynamics in the network during the

processing of string (1), “Betty get up.” Table E2.1 and all the followingtables list the word webs at the top and are referred to by the word they

represent. To the left and right of each word, the sequence sets connected

to its word web are indicated. To the left of each word, its past-orientedsequence sets are listed, and future-oriented sets appear to the right.

The left row of the table lists the time steps of the simulation. The second

column from the left gives the input to the grammar circuit – that is, the word

stimuli activating network elements. Global activity states are indicated in

the left column, also in capital letters.For each time step (rows) and neuronal set (columns), the table matrix

lists activity states. For each input unit, activity states are listed between

slashes. For each sequence set connected directly to an input unit, activity

states are listed either to the left (past sets) or the right (future sets) of the

slashes. The activity states introduced in Chapter 10 are distinguished – that

is, ignition ( I ), reverberation (R), priming (P ), and rest (0). Full ignition

Page 235: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 235/331

218 Basic Bits of Neuronal Grammar

Table E2.1. Time course of activation and deactivation of input sets (orword webs) and sequence sets of Fig. E2.1 during processing of the

string “Betty get up,” which is congruent with the network structure.Lines represent time steps. Time steps and input to the network arelisted on the left. In each column, activity states of an optionalpast-oriented sequence set (p; to the left of slashes), of a word web(between slashes), and of an optional future-oriented sequence set (f;to the right of slashes) are shown. Time steps between inputs arechosen so that activity spreads until stable states are reached. Lettersindicate ignition (I ), reverberation (R ), priming (P ), and rest (0). I 

indicates full unprimed ignition, and P and R denote the highest levels

of priming and reverberation.

  /Betty/f p/Get/f p/Up/

0 /0/0 0/0/0 0/0/

1 Betty /I∧/0 0/0/0 0/0/

2 /R/P 0/0/0 0/0/

3 /R/P P/0/0 0/0/

4 /R/P P/P/0 0/0/

5 gets /R/P P/I/0 0/0/

6 /R/P I/R/P 0/0/

7 /R/I R/R/P P/0/

8 /I/R R/R/P P/P/

9 /R/R R/R/P P/P/

10 up /R/R R/R/P P/I/

11 /R/R R/R/P I/R/

12 /R/R R/R/I R/R/

13 /R/R R/I/R R/R/

14 /R/R I/R/R R/R/

15 /R/I R/R/R R/R/

16 /I/R R/R/R R/R/

17 /R/R R/R/R R/R/

18 DEACTIVATE /0/0 0/0/0 0/0/

( I ∧) of a word web is distinguished from primed ignition ( I ). Reverberation

states are labeled by numbers indicating the rank of reverberation levelsin the activity hierarchy, R denoting the highest activity level, R

denoting

the next level below R, and R indicating the third highest level. Note that

these labels differ from those used earlier (e.g., R1, R2), a change made to

make the present tables easier to survey.

The time steps where input is given to the grammar network are chosen

arbitrarily, and to avoid cross-talk of activity that depends on exact timing.

Page 236: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 236/331

E2.2 Grammar Circuits at Work 219

The sentences should be accepted regardless of whether they are spoken

quickly or slowly, with or without pauses between words, and regardless of 

whether additional constituents are being inserted between any two of thewords. Therefore, the crucial dynamics in the network should not depend

on exact delays between word inputs. As a rule, several time steps in whichactivity is allowed to spread through the network lie between any two sub-

sequent inputs. The circuits are always allowed enough time to settle into

one of their stable states before a new input is computed.

Principles underlying activity dynamics have been explained in great de-

tail in Chapter 10 and are mentioned here only as far as they become rele-

vant in the individual derivations. The most important processes are asfollows:

   r Words in the input activate word webs (ignition I  if primed; full ignition I ∧ if not).

   r After its ignition, a set reverberates.   r Ignitions spread to adjacent sets that are already strongly primed.   r Reverberating word webs prime their future sets, which in turn prime

their adjacent sequence set and word web.   r Full ignition of a set reduces the activity state of all other sets that are in

the state of reverberation or priming.

E2.2.1 Simulation 1: Acceptance of a Congruent String

Table E2.1 presents activity dynamics caused in the grammar network bythe congruent sentence (18).

(18) Betty get up.

This sentence is assumed to be consistent with the network structure in

Figure E2.1. It can be considered to be grammatical in English if it is usedto give a command or recommendation.

Input units are being activated one after the other at time steps 1, 5, and

10, respectively [following principle (A2) in Section 4.5]. At the same time,

they ignite ( I ). After ignition of each input unit, its future-oriented sequence

detector is being primed [cf. principles (A3ii), (A3iii)]. Two time steps later,priming is further transmitted to the right, where it affects input units of pos-sible successor words (A3iv). This priming is important for later processing.

Crucially, it makes possible the primed ignition I  of the input unit activated

by the next input word. When primed (rather than full) ignition takes place,

the threshold regulation mechanism is not invoked. Thus, threshold regula-

tion only occurs when the first input unit ignites (A2i, A5i). At time step 1,

Page 237: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 237/331

220 Basic Bits of Neuronal Grammar

there are no preactivated sets at R1 or P 1; therefore, threshold regulation

has no effect.

After primed ignitions at time steps 5 and 10, waves of ignitions are ini-tiated that spread to the left, where sequencing units are in the states of 

priming or reverberation (A3i, A4ii, A4iii). These backward activity wavesare of utmost importance in the present network type because they establish

coherence between-word webs directly connected by sequence sets. By this

mechanism, the network establishes a syntactic match, the complement re-

lation, between two words. Backward spreading of ignitions occurs between

time steps 5 and 8, and, again, between time steps 10 and 16. The “back-

ward transmission” of activity synchronizes input units. The last backwardspreading wave establishes synchrony among all neuronal sets activated in

the present simulation. It resets the reverberation levels of all participating

units so that they can be assumed to reverberate in a coordinated and well-

timed manner, or if spreading of the ignition wave is assumed to happen

rather quickly, even almost synchronously.

Ignitions of the leftmost sequence unit in the diagram (representing thesentence initial proper name) at time steps 8 and 16 terminate the rightward

wave of ignitions. This is so because there are no remaining preactivatedunits to which ignitions could spread and refractoriness applies (A1v). This

deactivation indicates that all syntactic matching processes were successful.

Finally, note that the criteria for “string acceptance” are met. After time

step 16, all input units are active together with their set of sequencing units.Input units are therefore satisfied (criterion 1). All sets reverberate at the

highest level of R1. Therefore, they are called visible (criterion 2). Finally, a

coherent wave of ignitions has involved all of them. The input units are there-

fore synchronized (criterion 3). Because the criteria of satisfaction, visibility,and synchronization are reached, the string can be said to be accepted by thenetwork, and the active representations are finally being reset to 0. This

illustrates how the network accepts grammatical string (18), “Betty get up.”

E2.2.2 Simulation 2: Processing of an Incongruent String

What would be the network response to an ungrammatical or less regular

string? The word sequence (19) is not consistent with the network structureand processing of this string will therefore be used as an example of the

processing of an “ungrammatical” word sequence.

(19) Up get Betty.

Intuitively, the network may detect the mismatch between string and

network structure by assessing that words occurring in direct or indirect

Page 238: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 238/331

Page 239: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 239/331

222 Basic Bits of Neuronal Grammar

In this case, not a single criterion from the three criteria required for

string acceptance is met. First, because the “bridge” of primed sequence

detectors is missing after the ignitions, no synchronization can arise. Theactivity levels finally achieved are also different from each other, because

each element in the input elicited a full ignition, causing threshold regulation,which was in turn a result of the lack of priming. Consequent to the unprimed

ignitions, threshold control is activated (A5i) so that the excitation levels of 

the representations of earlier string elements are changed to lower levels

of priming. Therefore, two of the three input units finally reverberated at

lowered levels of R and were thus not visible. Furthermore, the three input

units ended up not satisfied – that is, they lack a full set of sequence detectorsat the highest reverberation level at the end of the computation processes.

A situation not covered by the comments in Chapter 10 arises in the

computations described in Table E2.2. At time steps 7 and 11, there are sets

that should exhibit two different activity states at a time. For example, there

is an input unit reverberating at R2 (or R) that, in addition, is primed at

a different level, P 1 (or P ). Dealing with this situation is discussed in moredetail in Chapter 12. In the tables in this Excursus, the level of reverberation

is taken into account and the additional priming is ignored.

E2.2.3 Simulation 3: Processing of a Partly Congruent String

A slightly different behavior of the same network is shown in Table E2.3.Here, string (20) is processed.

(20) Up Betty get.

Again, this string is inconsistent with the network structure and should be

taken as another example of an ungrammatical string. Although the network

is not prepared to accept it as a whole, word string (20) includes a substring

for which a sequence detector would be available. The ordered word pair

“Betty get” is the initial part of congruent sentence (18). String (20) shouldtherefore be partly acceptable for the circuit.

The misplaced initial element fails to cause priming of representations of 

subsequent elements. The ignition at time step 1 does not cause such primingof the neuronal set corresponding to the second word in the string. Therefore,

a full ignition  I ∧ occurs at time step 4. The reverberation level of the initiallyactivated set is lowered as a result of the activation of threshold control.

Because the activation processes following time step 4 finally lead to the

priming of the neuronal set representing the last element of the string in

the input, its presence in the input causes a primed ignition at time step 8,

followed by backward spreading of ignitions. This process can be considered

Page 240: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 240/331

E2.2 Grammar Circuits at Work 223

Table E2.3. Time course of activation and deactivation ofinput sets (or word webs) and sequence sets of Fig. E2.1 during

processing of the string “Up Betty get,” which is incongruentwith the network structure. For explanation, see Table E2.1. P ’and R ’ indicate reduced activity levels and P ” and R ” furtherdecreased activity states.

  /Betty/f p/Get/f p/Up/

0 /0/0 0/0/0 0/0/

1 Up /0/0 0/0/0 0/I∧/

2 /0/0 0/0/0 0/R/

3 /0/0 0/0/0 0/R/

4 Betty /I∧/0 0/0/0 0/R/

5 /R/P 0/0/0 0/R’/

6 /R/P P/0/0 0/R’/

7 /R/P P/P/0 0/R’/

8 gets /R/R P/I/0 0/R’/

9 /R/R I/R/P 0/R’/

10 /R/I R/R/P P/R’/

11 /I/R R/R/P P/R’/

12 /R/R R/R/P P/R’/

the network equivalent of a judgment that a substring of the stimulus is

coherent.

The backward wave of ignitions cannot spread to the neuronal set of the

first string element. Crucial for this is that there is, at time step 8, no priming

of the sequence detectors that could serve as a bridge between the relevant

input representations.Final synchronization of all word webs therefore fails to take place. In

this case, however, the synchronization and visibility conditions are met

for the representations of the last two string segments. Nevertheless, two of 

the three input units are not being satisfied and there is no synchronization

of representation of the well-formed string part and the misplaced particle.

Therefore, the network does not accept the string. Comparison of simula-

tions 2 and 3 shows that degrees of disagreement between network structure

and serial order of strings can be reflected in the sequence of activation anddeactivations in the grammar network.

These examples illustrate very elementary properties of a serial-order

circuit. To demonstrate that neuronal grammar circuits can process relevant

syntactic structures in a biologically realistic manner, it is necessary to con-

sider more complex and therefore more demanding examples. This is done

in Excursus E3.

Page 241: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 241/331

EXCURSUS THREE

 A Web Response to a Sentence

Simulationing sentence processing in grammar circuits is important because

it shows the processes that postulated by a neuronal grammar to occur in

the brain when congruent and incongruent word and morpheme strings are

processed. The examples discussed in Chapter 10 and Excursus E2 were in-troduced to illustrate the working of the envisaged grammar machinery, the

principles of which are less obvious from a more complex simulation. How-ever, the earlier examples can be considered to be toy simulations because

the strings under processing exhibit far less complexity than most sentences

commonly used in everyday language.

It is therefore relevant to look at more complex examples of neuronal cir-cuits that may be the basis of syntactic knowledge and syntactic processing.

In this Excursus, a sentence discussed earlier in the context of conventional

grammar models (see Chapter 7) is again the target. First, the algorithmic ver-

sion of a neuronal grammar processing this and similar sentences is presentedand the corresponding network described. Subsequently, activity dynamics

caused in the grammar circuit by the sentence in the input are discussed.An animation of this simulation is available on the Internet at the books

accompanying web page (http://www.cambridge.org).

We look first at sentence (1).

(1) Betty switches the machine on.

Putative syntactic structures possibly underlying the processing of sentence(1) are spelled out in the context of dependency and rewriting grammars

(Section 7.5; Fig. 7.1). Also, a very tentative neurobiological circuit was pro-

posed earlier that may underlie the processing of (1). This earlier first-order

approximation at a grammar circuit must now be revised in the light of con-

clusions drawn in Chapters 10 and 11.

224

Page 242: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 242/331

E3.1 The Grammar Algorithm and Network 225

In contrast to the toy examples used in Excursus E2, sentence (1) exhibits

nontrivial features, including the following:

   r

Discontinuous constituent made up by the verb and its particle   r Agreement between subject and verb suffix   r “Left branching” of the verb phrase.

As explained in Chapter 7, distributed words such as switch ...on – called particle verbs here because the verb root requires a particle (“on”) – are dif-

ficult to capture by syntactic tree structures. Transformations, feature trans-

port through the branches of trees or other additional mechanisms must be

assumed to make the relationship between the two distant lexical elementsmanifest in the syntactic representation. The same problem arises for theagreement between the proper name and the verb affix; again, two spatially

or temporally separate elements (the proper name Betty and the verb suf-

fix–(e) s) must be linked. Left branching refers to the fact that in a standard

phrase structure representation of sentence (1) as depicted in Figure 7.1, the

VP(orV) node carries two downward branches, one of which (the left one)branches again. Multiple branches to the left have been proposed to give

rise to computational problems. It is relevant to confront neuronal grammarnetworks with strings including discontinuous constituents, agreement, and

left branching, because these features can pose problems to grammars and

sentence processors.

Instead of postulating different syntactic mechanisms to capture rela-tionships between adjacent and distant string elements, as most grammar

theories do, neuronal grammar bases short- and long-distance dependencies

on a single mechanism provided by sequence sets (cf. Chapters 9 and 10).

Constructions necessitating leftward branches in the context of rewritinggrammars may be analyzed differently within other grammar frameworks

(cf. Fig. 7.1). The “flat” representations of neuronal grammar, in which allcomplement lexical items are each linked directly to the other, avoids the

assumption of leftward branching, at least for many sentence types in which

a conventional grammar may postulate such.

E3.1 The Grammar Algorithm and Network

As stressed earlier, depicting a grammar fragment as a network is not the

only way of representing it. The network can also be rewritten as an equiva-

lent set of formulas in which lines indicate connections between word webs

and neuronal sets processing information in the past or future that make

up lexical category representations. A neuronal grammar fragment that can

process (1) is given next. Three sets of formulas – assignment, valence, and

Page 243: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 243/331

226 A Web Response to a Sentence

sequence formulas – make up the algorithmic version of the grammar frag-

ment. For explanation of terms, see Chapter 7, and for an explanation of the

formulas, see Chapter 11.The following assignment formulas specify a set of connections between

representations of morphemes and their respective lexical category repre-sentations composed of sequence sets. Lexical categories are labeled by

abbreviations, and their full names are given in brackets.

(2) Betty ↔ Prop (nominative proper name)

(3) switch ↔ V (transitive particle verb)

(4) -es ↔ Vs (verb suffix)

(5) the ↔ Det (determiner)(6) machine ↔ N (accusative noun)

(7) on ↔ Vpart (verb particle)

When extending the grammar to a larger vocabulary, it is necessary to clar-

ify these categories, some of which can be defined rather widely. All proper

names, for example, should be part of the proper name category. The cat-

egory of proper names would be closely related to that of ordinary nomi-

native nouns. The distinctive feature of the two would be the past-orientedsequence detector of the ordinary noun category that responds to a pre-

ceding determiner, which is usually present with nouns that are not proper

names. Note that the distinction between proper and other names has syn-

tactic implications, but can be defined on the basis of semantic criteria as

well.In contrast to very widely defined categories with many members, the

verb suffix category has only a few members. These members would be the

different realizations of the verb suffix signaling present tense, third person,and singular. In written English, this suffix can be s or es, and in spoken

language, the s can be voiced or unvoiced and further context variants of the

acoustic pattern of the suffix can also be distinguished. These variants of the

morpheme could be realized in the model as overlapping sets, the abstractrepresentation of the third-person singular present suffix being represented

by the intersecting parts of these sets. (A similar proposal was made at the

semantic level in Section 5.2.3 to model family resemblance.) Vpart would

also be a very much restricted “lexical category.” It would only include oneverb particle because choice of a different particle would imply a different

meaning and grammatical implication of the particle verb.The grammar fragment includes a set of lexical category representations

that can be defined as sets of sequence sets. The following valence formulas

define the lexical categories in terms of the sequence sets that constitute

them.

Page 244: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 244/331

E3.2 Sentence Processing in Syntactic Circuits 227

(8) Prop (/∗/ f 1, f 2)

(9) N (p1, p2 /∗/)

(10) V (p /∗

/ f 1, f 2, f 3)(11) Vs (p1, p2 /∗/)

(12) Det (/∗/ f)

(13) Vpart (p /∗/)

Each letter p (past set) refers to a sequence set sensitive to words or mor-

pheme from the lexical category preceded by words of a different category,

and each f (future set) refers to a sequence detector sensitive to a word ormorpheme of that category followed by an element of a different category.

Indices are used if more than one past- or future-sensitive sequence set ispresent in any given lexical category set.

The reciprocal but asymmetric connections between all sequence sets are

specified by the following set of sequence formulas:

(14) Prop (f 1)→→  V (p)

(15) Prop (f 2)→→  Vs (p2)

(16) V (f 1) →→  Vs (p1)

(17) V (f 2) →→ N (p2)(18) V (f 3) →→  Vpart (p)

(19) Det (f) →→ N (p1)

Note that all past- and future-oriented sequence sets are connected. The

sequence and valence formulas taken together include the information that

lexical category representations are connected to each other directly.The network equivalent to these formulas is presented in Figure E3.1.

Each small circle represents a neuronal set. Word webs, the neuronal unitsprocessing words or morphemes, are listed in the top line with the word or

affix they are specialized for indicated. Below each word web, linked directly

by arrows, are the sequence detectors the word web is connected to. The

set of sequence detectors connected directly to a word web represents thelexical category representation the word is part of. These direct connections

betweenwordwebandsequencesetsarespecifiedbyvalenceandassignment

formulas. The horizontal connections between sequence sets correspond to

sequence formulas (14)–(19).

E3.2 Sentence Processing in Syntactic Circuits

Table E3.1 displays activity dynamics of the network processing sentence

(1). The reader is now invited to follow the activity waves through the net-

work. The principles underlying the dynamics are the same as in earlier

Page 245: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 245/331

228 A Web Response to a Sentence

Figure E3.1. Fragment of neuronal grammar that processes and accepts the sentence“Betty switches the machine on.” (An equivalent formalism is provided in the text.)Activation and deactivation dynamics of the network are shown in Table E3.1.

simulations (cf. Chapter 10 and Excursus E2). First, the global characteris-

tics of the activation and deactivation processes in this simulation are briefly

summarized. Finally, a few special features of this simulation are stressed.

E3.2.1 Global Characteristics

All activation and deactivation processes are described in a table organizedinthesamewayasearlieractivationtablesofthiskind.Foradetaileddescrip-

tion of how activation tables represent spreading activity through a network,

see Section E2.2 in Excursus E2. The principles of neuronal grammar thatyield the activation and deactivation processes have been discussed in great

detail in Chapter 10 (e.g., in Section 10.7).

Each lexical element in the input activates its input unit. Here, the verb

suffix – the third-person singular present suffix realized, in this case, as

-es – is represented as a separate lexical element. Therefore, the five-wordsentence activates six word or morpheme webs in a given order. The timesteps at which word webs ignite are 1, 5, 10, 16, 20, and 29. As explained

earlier, when a new item is presented, the time steps are chosen such that

the network settles into a stable state before a new input is given. Each word

or morpheme in the input causes an ignition of the word web representing

it. Except for the first word, Betty, and the article the, the ignitions caused

Page 246: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 246/331

E3.2 Sentence Processing in Syntactic Circuits 229

are primed ignitions, I , because the respective input units had already been

preactivated. The other two ignitions are full ignitions, abbreviated as I ∧.

The ignitions change the activity states of other sets that are connecteddirectly to them. The ignition itselfand the reverberation state following each

ignition produce priming of the directly connected sequence set sensitiveto information about future events. Spreading of priming is indicated by

changes of the inactive state, 0, to the state of priming, P . In Table E3.1, these

changes are indicated to the right of word webs ignited by external input.

Priming of strongly connected sets takes place after each ignition caused by

input, except for the last one (time step 29). Priming can be effective up to

the next word web connected to the reverberating input unit by sequencesets ( future sets). Another way to express this would be to say that a word

prepares the circuit to perceive a word from a category whose members

frequently follow the word. However, if the next word web has more than

one connected sequence set sensitive to information available in the past

(labeled p plus index), all of its past sets need to be primed in order to cause

priming of the word web (cf. Section 10.7). Ignitions also momentarily primetheir directly connected sequence sets to the left. This temporary effect is

not covered by Table E3.1, except for in the cases in which it has long-lastingconsequences (traveling waves of ignitions).

An ignition caused by external input – the word or morpheme stimuli –

leadstoadditionalprimedignitionsindirectlyconnectedsetsthatarealready

in a preactivated state (that is, sets that already reverberate or are in the stateof priming at high levels). This leads to backward-traveling waves of activity,

from right to left in the table, which are said to synchronize the sets involved.

Backward activity spreading of primed ignitions occurs in the intervals of 

time steps 5–8, 10–14, 20–27, and 29–38.There is also the possibility of indirect influence of an ignition on other

neuronal sets via activation of threshold regulation processes. This becomeseffective only after the determiner the has been presented, following time

step 15. Activity levels of all preactivated sets in the network are reduced,

from R1 to the lower level R2, or from P 1 to P 2, as a consequence. As

in Excursus E2, the canonical activity levels R1, R2, R3 . . . and P 1, P 2,P 3 , . . . are given as R, R, R, . . . and P , P , P  in the table.

After the sentence had occurred in the input, activity reaches a stable stateat stime step 39. At this point in time, the network has all its word webs or in-

put units reverberating at the highest reverberation level R (also called R1).

The visibility condition is therefore satisfied. In addition, all of the word

webs have highly active connected lexical category representations (i.e., sets

of sequence sets. This means that the condition of satisfaction is also met.Finally, a terminal wave of ignitions spreading throughout all activated sets

Page 247: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 247/331

Table E3.1. Time course of activation and deactivation of word webs andsequence sets of Figure E3.1 during processing of the string “Bettyswitches the machine on,” which is congruent with the network structure.Lines represent time steps. Time steps and input to the network are listed

on the left. In each column, activity states of one or more past-orientedsequence sets (p; to the left of slashes), of a word web (between slashes),and of future-oriented sequence sets (f; to the right of slashes) are shown.Time steps between inputs are chosen so that activity spreads until stablestates are reached. Content words are abbreviated (B. – Betty, s. – switch,m. – machine). Letters indicate ignition (I ), reverberation (R ), priming (P ),and rest (0). I ∧ indicates full unprimed,ignition. P and R denote thehighest levels of priming and reverberation and P 

and R  indicate lower

activity levels.

 /B./f 1,f 2 p/s./f 1,f 2,f 3 p1,p2  /-s/ /the/f p1,p2  /m./ p/on

1 Betty /I∧/0,0 0/0/0,0,0 0,0/0/ /0/0 0,0/0/ 0/0/

2 /R/P,P 0/0/0,0,0 0,0/0/ /0/0 0,0/0/ 0/0/

3 /R/P,P P/0/0,0,0 P,0/0/ /0/0 0,0/0/ 0/0/

4 /R/P,P P/P/0,0,0 P,0/0/ /0/0 0,0/0/ 0/0/

5 switch /R/P,P P/I/0,0,0 P,0/0/ /0/0 0,0/0/ 0/0/

6 /R/P,P I/R/P,P,P P,P/0/ /0/0 0,0/0/ 0/0/

7 /R/I,P R/R/P,P,P P,P/P/ /0/0 P,0/0/ P/0/

8 /I/R,P R/R/P,P,P P,P/P/ /0/0 P,0/0/ P/P/

9 /R/R,P R/R/P,P,P P,P/P/ /0/0 P,0/0/ P/P/

10 es /R/R,P R/R/P,P,P P,P/I/ /0/0 P,0/0/ P/P/11 /R/R,P R/R/P,P,P I,I/R/ /0/0 P,0/0/ P/P/

12 /R/R,I R/R/I,P,P R,R/R/ /0/0 P,0/0/ P/P/

13 /I/R,R R/I/R,P,P R,R/R/ /0/0 P,0/0/ P/P/

14 /R/I,R I/R/R,P,P R,R/R/ /0/0 P,0/0/ P/P/

15 /R/R,R R/R/R,P,P R,R/R/ /0/0 P,0/0/ P/P/

16 the /R’/R’,R’ R’/R’/R’,P’,P’ R’,R’/R’/ /I∧/0 P’,0/0/ P’/P’/

17 /R’/R’,R’ R’/R’/R’,P’,P’ R’,R’/R’/ /R/P P’,0/0/ P’/P’/

18 /R’/R’,R’ R’/R’/R’,P’,P’ R’,R’/R’/ /R/P P’,P/0/ P’/P’/

19 /R’/R’,R’ R’/R’/R’,P’,P’ R’,R’/R’/ /R/P P’,P/P/ P’/P’/

20 machine /R’/R’,R’ R’/R’/R’,P’,P’ R’,R’/R’/ /R/P P’,P/I/ P’/P’/

21 /R’/R’,R’ R’/R’/R’,P’,P’ R’,R’/R’/ /R/P I,I/R/ P’/P’/

22 /R’/R’,R’ R’/R’/R’,I,P’ R’,R’/R’/ /R/I R,R/R/ P’/P’/

23 /R’/R’,R’ R’/I/R’,R,P’ R’,R’/R’/ /I/R R,R/R/ P’/P’/

24 /R’/R’,R’ I/R/I,R,P R’,R’/R’/ /R/R R,R/R/ P’/P’/

25 /R’/I,R’ R/R/R,R,P R’,I/R’/ /R/R R,R/R/ P/P’/

26 /I/R,R’ R/R/R,R,P R’,R/I/ /R/R R,R/R/ P/P/

27 /R/R,I R/R/R,R,P I,R/R/ /R/R R,R/R/ P/P/

28 /R/R,R R/R/R,R,P R,R/R/ /R/R R,R/R/ P/P/

29 on /R/R,R R/R/R,R,P R,R/R/ /R/R R,R/R/ P/I/

30 /R/R,R R/R/R,R,P R,R/R/ /R/R R,R/R/ I/R/

31 /R/R,R R/R/R,R,I R,R/R/ /R/R R,R/R/ R/R/32 /R/R,R R/I/R,R,R R,R/R/ /R/R R,R/R/ R/R/

33 /R/R,R I/R/I,I,R R,R/R/ /R/R R,R/R/ R/R/

34 /R/I,R R/R/R,R,R R,I/R/ /R/R I,R/R/ R/R/

35 /I/R,R R/R/R,R,R R,R/I/ /R/R R,R/I/ R/R/

36 /R/R,I R/R/R,R,R I,R/R/ /R/R R,I/R/ R/R/

37 /R/R,R R/R/R,R,R R,R/R/ /R/I R,R/R/ R/R/

38 /R/R,R R/R/R,R,R R,R/R/ /I/R R,R/R/ R/R/

39 /R/R,R R/R/R,R,R R,R/R/ /R/R R,R/R/ R/R/

40 DEACTIVATE /0/0,0 0/0/0,0,0 0,0/0/ /0/0 0,0/0/ 0/0/

Page 248: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 248/331

E3.2 Sentence Processing in Syntactic Circuits 231

(time steps 29–38) follows the last stimulation by an element in the input.

The synchronization condition is therefore met as well. The sentence is in

accord with the structure of the circuit. Therefore, the activations inducedby the sentence are being terminated by the postulated deactivation process.

E3.2.2 Specific Features

It may be worthwhile to consider the following special features of the pro-

cessing of sentence (1) that address the three specific features of its syntactic

structure – namely, the included discontinuous constituent, the agreement

exhibited by two of its constituents, and the left branching attributed to itby a phrase structure grammar (PSG).

The discontinuous constituent switch . . . on does not pose problems to

this processing device because the particle is bound to the verb by the same

kind of connection assumed between the verb and its other complements.

This is related to the fact that in a neuronal grammar, there is no motivationto prohibit what would be crossings of projection lines in a PSG. Crossings

of projection lines are prohibited in rewriting systems, although there is

probably no biological motivation for this. Following time step 5 of thissimulation, the verb stem primes its connected sequence detectors that also

link it to a verb suffix potentially activated in the future. Priming of the verb

suffix representation becomes effective at time step 8, so that its activationat time step 29 is primed ignition.

Also, the fact that choice of the verb suffix depends on both the subject

and the verb is transformed naturally into the network. Connections are

postulated from both the proper name Prop to the verb suffix Vs, and, in

the very same way, from the verb V to the verb suffix Vs. Neuronal gram-mar treats agreement between verb and verb suffix in the same way as anyother dependency (cf. Chapter 9). Again, the point here is that dependency

relations need not be two dimensional, but can have as many dimensions

as required to describe cooccurrence of morphemes in sentences. In this

simulation, the process most important in this context – the priming of the

verb suffix representation by two sequence sets, one of them sensitive to the

earlier proper name and the other to the preceding verb – becomes effective

at time step 7.The determiner the preceding the noun machine does not have a primed

neuronal representation when this word occurs in the input. This is because

the article’s web linked only to the noun’s web, and the noun web has not yet

been activated. A full, rather than primed, ignition is therefore achieved by

the determiner that reduces activity states of already active representations.There is a danger that the already active and the newly activated represen-

tations will fail to be linked by synchronization processes, a situation that

Page 249: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 249/331

Page 250: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 250/331

E3.3 Discussion of the Implementation and Simulation 233

One perspective for further development is offered by the introduction

of connections of different strengths. Only one type of strong (forward)

and weak (backward) connection has been set up in the grammar networksdiscussed in detail here. The present proposal could be further developed to

capture the fact that the verb suffix must occur directly after the verb stem,whereas accusative noun phrase and particle can be placed later in the string.

In principle, differential connection strength could realize this. Connection

strength 1.0 could be postulated for the link of V to Vs, but a weaker link,

0.8, could be assumed between V and Vpart. The stronger links could then

be effective earlier and the weaker ones only later. This would not remove

the advantage of allowing for free placement of complements [cf. sentences(1) and (20)] because connection strengths between a constituent and twoof its complements can always have the same value, thus leaving open which

constituent must occur first. Because differential connection strengths would

make the present computations more complex, this possibility is not dealt

with here in more detail.

One may argue that it might be a caveat of the approach proposed here

that not only the standard word order is possible in English, and this prob-

lem is even more virulent for other languages. The simple solution suggested,however, is that different grammatical word orders of the same or similar

sets of constituents can be processed by different sets of sequence feature

detectors. An active and a passive sentence and a main clause and a subor-

dinate clause would therefore be characterized by distinct sets of sequencedetectors. As argued in Section 10.6.4, it is not costly to add sequence sets

for processing strings with different word order.

One may argue that the philosophy underlying this kind of sentence pro-

cessing has some relationship to standard grammatical theories and thatthere is nothing in these computations that a modern grammar could not do.

This is clearly correct. By introducing concepts such as slash categories, oroperations such as movement or tree adjoining, the capacity of a grammar

algorithms have been greatly augmented, so that the claim cannot be made

that neuronal mechanisms cover strings that conventional grammars fail to

process.

A major difference between conventional syntax theories and neuronal

grammar is, of course, that the present proposal attempts to specify brainmechanism on a concrete level. A further difference to a many standard

grammatical analyses lies in the fact that these grammars compute a tree

representation as an output, whereas the neuronal automaton has a group

of satisfied, visible, and synchronized neuronal sets activated and finally de-

activated. It is possible that this is, at a certain level of abstraction, equivalent

to one or more nonneuronal grammar algorithms. Chapter 11 claims that a

Page 251: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 251/331

234 A Web Response to a Sentence

certain modification of dependency syntax is equivalent to grammar circuits.

It is not impossible that slight modification of other grammar types lead to

descriptions that are compatible with the present concept of a neuronal

grammar.

A further criticism of neuronal grammar could be as follows: It could be

argued that the processes in grammar networks lead only to the activation

and grouping of representations of lexical items. This could be considered

to be similar to listing the lexical items included in a sentence together with

one bit of information about whether the string was grammatical. It may

be claimed that too much information about sentence structure and hierar-

chical relationships between lexical elements would be lost if computationswere preformed in this manner. The argument may therefore be that a tree

representation is preferrable because it preserves more structural properties

of the morpheme string – in particular, its hierarchy.

However, this putative critique does not appear to be appropriate. In

grammar circuits, word representations become active together with one of 

their respective lexical category representations and their neuronal links.

Together with the activity dynamics of neuronal sets, these provide a struc-

tural characterization of the perceived string in much the same way as atree graph, but with the important difference of not being restricted to two-

dimensional representations. Furthermore, additional information becomes

available in the course of sentence processing. During the computations in

the circuits, information about relationships between lexical items is revealed

by the process of synchronization and the level of reverberation of the sets.

From the computation summarized in Table E3.1, it is clear that the first

three elements of the sentence belong together from the beginning because

they exhibit synchronization after the verb suffix leads to ignition of its inputunit. Also, the close relationship between noun and article becomes visible

in the adjacent ignitions of their input and sequencing units. In the final

synchronization process initiated by the last word in the input, information

about dependency relations between lexical items is revealed by the order in

which their representations ignite. Thus, it appears that the present frame-

work is quite capable of revealing information about dependency relations

in sentence structures, although the tree structure hierarchies are usually not

assumed in this type of grammar approach.An additional advantage of this approach is that it offers a clear-cut

description of the sequence of activation and deactivation processes, the

sequence of groupings of constituents, and the temporal order in which

synchrony is being achieved in the course of the processing of a sentence.

Structural information about a sentence therefore unfolds over time in the

grammar circuit.

Page 252: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 252/331

CHAPTER TWELVE

Refining Neuronal Grammar

This chapter takes the concept of a neuronal grammar developed earlier as

a starting point. The earlier proposal is partly revised and extended to cope

with problems put forth by natural language phenomena.

The three phenomena considered here are as follows:

   r The distinction between a constituent’s obligatory complements and its

optional adjuncts.   r The multiple occurrence of the same word form in a string.   r The embedding of sentences into other sentences.

All three issues have been addressed occasionally before; however, they were

not treated systematically in the context of neuronal grammar. This requires

its own discussion because the necessary extensions lead to a substantial

revision of the previous concept of neuronal grammar.

The gist of the revision briefly in advance is as follows: In earlier chapters,the relationship between sequence detectors and words in the input was

assumed to be static. The word web ignites and then, successively, one set

of its sequence detectors is recruited according to relationships the word

exhibits to words in its context, as they are manifest in regular co-occurrence

of lexical category members. Each word would recruit one of its connected

lexical category representations. However, if each word form were attributed

to one lexical category, it would be impossible to model the situation in which

one word occurs twice in different grammatical roles in a sentence. In thiscase, the single word representation must store the knowledge about the

two occurrences, and the knowledge about the syntactic role the particular

word occurrences played in their respective contexts as well. Section 12.2

discusses a preliminary solution to this dilemma. The idea here is that each

word representation can exhibit several different activation states at a time.

Two different waves of activity spreading through the same neuronal set may

235

Page 253: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 253/331

236 Refining Neuronal Grammar

allow for storing the fact that a given item was present twice. The proposed

mechanism can even operate if a lexically ambiguous word is used twice in

a sentence, as member of two different lexical categories.An even more radical use of the neuronal grammar machinery is offered

in Section 12.3, where, finally, center embeddings are addressed. The solutionexplored here is that links between word webs and sequence sets are fully

dynamic. The envisaged neuronal basis of the dynamic link is the traveling

of a coherent wave of activity that reverberates back and forth between

neuronal sets.

At the start, an important linguistic distinction between complements and

adjuncts is reviewed and a few thoughts about its neuronal basis are added(Section 12.1).

12.1 Complements and Adjuncts

In Chapter 10, lexical elements were considered as units that “require” oth-

ers. If one lexical element is present in a string, its complements can also beexpected, with the converse also true – the “dependent” item also requires

the “governor.” This view gave rise to the concept of mutual dependence thatavoids the hierarchy implication. Nominative noun and verb are mutually

dependent. This is implemented by a reciprocal connection, with stronger

links in the direction of the normal succession of the elements (from nomi-

native noun to verb in English) than in the reverse direction. The postulateof reciprocal connections between neuronal sets α and β is based on the

following assumptions:

   r

Large cortical neuron populations are connected to each other recipro-cally.

   r Strong correlation characterizes the occurrence of the words or mor-phemes represented by sets α and β.

These issues are addressed in earlier chapters (e.g., Chapter 2, in which the

reciprocity assumption is discussed; Chapter 10, in which mutual dependence

is based on reciprocal asymmetric connections).

In contrast to the case in which two types of lexical elements are correlatedand usually cooccur, there are cases in which the occurrence of one lexicalelement depends on the occurrence of the other, but the reverse is not true.

Whereas a noun (or proper name) and a verb usually co-occur, as in (1),

the relation between a relative pronoun and a noun/proper name should

probably be characterized as follows: If one occurs, the other is usually also

present, but the converse is not necessarily so.

Page 254: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 254/331

12.1 Complements and Adjuncts 237

(1) Hans comes . . . .

(2) Hans who . . . .

The relative pronoun who can be added freely to any proper name referring

to a human being, but the relative pronoun necessitates the noun or a similar

element to which it can adjoin. If there is no noun or related item, no relative

pronoun can occur. The nonobligatory, to be a freely added element is calleda free adjunct . The distinction between obligatory complements and free

adjuncts is common in grammar theories (see Ajukiewicz, 1936; Haegeman,

1991; Hays, 1964; Tesniere, 1953).

Note again the asymmetry between the adjunct and the noun: The wordwho used as a relative pronoun requires the occurrence of a noun or some-

thing similar (e.g., proper name, pronoun). The probability of finding who

after a noun is generally low, but if a relative pronoun who is present, it is

almost certain that also a noun or noun-like element (e.g., personal pronoun,

proper name) is in the string.

How would this characteristic asymmetry of adjuncts be realized in a

neuron circuit? Would it be realized through the same type of reciprocal

and directed connection as postulated for the links between mutual comple-ments?

It is well established that coactivation of a pre- and postsynaptic neuron

leads to synaptic strengthening (see Chapter 2). However, presynaptic acti-

vation and postsynaptic silence can reduce synaptic weights dramatically.

This process is called homosynaptic long-term depression (see Tsumoto,

1992). The reverse situation can have a similar effect as well. In this case,presynaptic silence while the postsynaptic neuron fires also reduces weights,

a process called heterosynaptic long-term depression. Both processes, ho-mosynaptic and heterosynaptic long-term depression, can lead to a reduc-

tion of synaptic weights between two connected neurons. It is still under

discussion which of the two processes produces stronger synaptic weight re-

duction. However, there are theoretical arguments why, in a network madeup of distributed and strongly connected sets of neurons, one of the pro-

cesses should have a stronger effect than the other. From computer simula-

tions, it became apparent that for achieving stable functioning of a system

of cell assemblies, it was necessary to introduce a “post- not pre-” heterosy-naptic rule of synaptic weight reduction (Hetherington & Shapiro, 1993).

One may therefore propose that presynaptic silence with postsynaptic firingleads to the most pronounced reduction of synaptic weights. The assump-

tion that long-term depression is more effective if it is heterosynaptic (post-

not pre-) gives rise to a prediction on the neuronal implementation of free

adjuncts.

Page 255: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 255/331

238 Refining Neuronal Grammar

If a neuronal element β is always active when α is active, but there are

many cases in which α is active while β is silent, then the connections from α

toβ

are assumed to suffer more from synaptic weight reduction than thosebackward from β to α. This postulate is based on a “pre- not post-” rule of 

weakening of synaptic connections. We assume here that the situation of alexical element of type b relying on the presence of another one of type a – as

the relative pronoun relies on the noun, or the who relies on Hans in example

(2) – is modeled in the network by unidirectional connections from neuronal

sets β to α. The idea here is that, although reciprocal connections between

sets can be assumed to be a default (see discussion in Chapter 2), learning

that a can appear without b, whereas b most frequently occurs together witha, causes one of the two links to become ineffective.

The unidirectional links imply that ignitions can spread only from the

adjunct representation to the “mother”, but no significant amount of activity

can be transmitted in the opposite direction, from the mother representation

to the adjunct. The adjunct–complement distinction is an important one for

language functioning, and its putative neurobiological realization outlinedhere plays a role in the syntactic circuits elaborated in this chapter and the

subsequent excursuses.

12.2 Multiple Activity States of a Neuronal Set

Multiple occurrence of lexical items and other syntactic objects is a genuine

characteristic of natural language. To allow words to occur repeatedly in a

string is at the heart of the view on language proposed by most mainstream

linguistic theories because, from this perspective, each language is consid-

ered to be a system that is enumerable recursively. This implies that rulescan be applied repeatedly in the analysis or synthesis of a sentence, and italso implies that, in principle, the number of possible sentences is unlimited.

One may want to call this the linguistic infinity position or linguistic infinity

view.

The linguistic infinity view implies that sentences are, in principle, allowed

to have infinite length. Also, because the base vocabulary of each language is

considered to be finite, this view further implies that lexical items are allowed

to occur repeatedly within sentences, with no upper limit, in this case, for thenumber of repetitions. Therefore, from this linguistic infinity perspective, it

appears imperative to specify possible mechanisms underlying processing of 

multiple occurrences of morphemes. From a more realistic viewpoint, it also

appears advantageous to have a mechanism at hand that allows one to have

multiple occurrences of the same lexical element in well-formed strings.

Page 256: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 256/331

Page 257: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 257/331

240 Refining Neuronal Grammar

After its ignition, a neuronal set is proposed to reverberate at R1. One

way to model reverberation is by using the paradigm of a huge neuron loop

in which not individual neurons but groups of neurons are connected insuch a way that the groups form a circle, or a more complex circuit includ-

ing many loops (Chapter 8). If there are directed connections between theneuron groups, an active group activates the next group that, in turn, acti-

vates the next, and so on. Circulation and reverberation of activity results

in such a model of reverberatory synfire chains. Important assumptions are

that reverberations last for tens of seconds, and that two or more separate

reverberatory waves of activity in the same network do not interfere with

each other.A few considerations on putative mechanisms underlying multiple rever-

beration may appear relevant here. A neuronal set has been proposed to

have an ordered inner structure, in the manner of a reverberatory synfire

chain (Chapters 8 and 10). A reverberating wave of activity in a neuronal

set may emerge after full activation of all neurons of the set because after

the ignition, refractory processes in strongly active neurons reduce the levelof activity, ultimately leaving only a small subset of neurons active. If an

igniting assembly includes a very strongly connected subset of neurons, itskernel (Braitenberg, 1978a), these neurons are the first to become strongly

active, and therefore fatigue (e.g., refractoriness) is first present in this sub-

set. Shortly thereafter, the kernel neurons are the first to become active

again, because their refractoriness is already over at the time when mostother set neurons are still refractory. Therefore, reverberations should start

in the kernel and spread from there through the rest of the set. This would

be the situation in a neuronal set activated only once.

The putative mechanism of multiple reverberation is the following. If awave of activity already reverberates in the set, a second ignition as a result

of external stimulation does not disturb the first wave, but causes a seconddistinct wave that also reverberates in the set at the same frequency, but

at a different phase. If, at the time of the second ignition, the set is in the

reverberating state, one group of neurons must be active and another set –

the subpart of the synfire chain that was active just one time step before –

must be in its refractory period. Therefore, it cannot take part in the second

ignition. For this reason, there is a group of neurons in the set that does notparticipate in the second ignition. These neurons in the refractory group are

among the first that become active again after ignition ceases. If an ignition

usually results in one wave of reverberating activity starting in the kernel, in

this case there are two such waves that emerge after the second ignition of 

an already reverberating set. Although wave one would be initiated in the

Page 258: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 258/331

12.2 Multiple Activity States of a Neuronal Set 241

kernel, the second wave would, so to speak, result from the “footprint” of 

the earlier reverberation.

It may be fruitful to explore these putative mechanisms further using com-puter simulations of reverberatory synfire chains (Bienenstock, 1996). Such

simulation can further explore the idea that, if a neuronal set is conceptual-ized as a reverberatory synfire chain in which well-timed waves of activity

resonate (see Chapter 8), more than one wave can reverberate at a given

point in time within any given set. The application of this idea to grammar cir-

cuits is explored further in Excursuses E4 and E5, in which more simulations

of the processing of sentences is discussed. In the following considerations,

it is assumed that input units – word and morpheme webs – can exhibit mul-tiple states of reverberation as a result of multiple stimulation by external

input. Multiple activity states can also be postulated for sequence sets.

12.2.2 Some Revised Principles of Neuronal Grammar

A neuronal set  can now be redefined as an entity that can exhibit several

simultaneous states of reverberation, priming, and ignition at any point in

time. An ignition induced by external input causes a wave of reverberationwithin the input unit and the number of ignitions during a given period

determines the number of its reverberations. It is emphasized that there is,

at present, no evidence available supporting this postulate. This postulate –oranalternativebasisofanaccountoftheprocessingofmultipleoccurrences

of the same language element – is necessary in any approach to grammar

mechanisms in the brain.

As a consequence, the ignition of an already reverberating input unit

should not result in a primed ignition but in a new full ignition instead.This makes revision of the principles outlined in Section 10.7 unavoidable.Furthermore, it is not clear how the idea of multiple reverberations of input

units matches that of threshold regulation as specified by axiom (A5) in

Section 10.7. If full (unprimed) ignition I ∧ reduces Ri to Ri+1, an additional

 I ∧ further diminishes all already established reverberation to Ri+2. Whereas

an input unit that is already in the state of reverberation usually exhibits

unprimed full ignition, it still exhibits primed ignition if being in a state of 

priming at a high level.To exclude overly strong activity because of multiple reverberation, a set

is only allowed one wave of reverberation at the highest level R1 at any

given time. If an ignition takes place while the set reverberates at R1, two

waves of activity emerge at R1 and R2, respectively. Next, input units areassumed to produce multiple reverberation as a consequence of ignition

Page 259: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 259/331

242 Refining Neuronal Grammar

by external input if they have already been in a state of reverberation Ri

before the ignition. Axiom (A2) must now be modified and reformulated

as (A2

):(A2’) Activity change in an input unit S  caused by external stimulation

(i) S  (0,t ), S  (E ,t ) ⇒ S  (I ∧,t +1)

(ii) S  (P 1,t ), S  (E ,t ) ⇒ S  (I ,t +1)

(iii) S  (P i ,t ), S  (E ,t ) ⇒ S  (I ∧,t +1)

if i  > 1

(iv) S  (R 1,t ), S  (E ,t ) ⇒ S  (I R 2,t +1)

(v)S 

(R 

i ,t 

),S 

(E 

,t 

)⇒ S 

(I ∧ R 

i ,t +

1)if i  > 1

where S ( x1, x2, . . . xh, y) is a predicate with h+ 1 places indicating the h

activity states (h > 0) of neuronal set S at time y. x1 can take the values 0, I ,

or any level of R or P , and x2 to x1 can only take levels Ri or P i , i > 1. Thus,

multiple activity states can be expressed, for example, by (6).

(6) S (I R 3 R 7, t )

This means that at time t , set S simultaneously ignites and reverberates atR3 and R7.

There is, in principle, no upper limit for the number h of simultaneous

activity states. If axioms (A2iv) or (A2v) are allowed to be applied h timesto the same input unit, the unit exhibits h activity states simultaneously.

Although an abstract version of the model may postulate high numbers of h, a realistic model would probably limit the number of each word web’s

simultaneous activity states to a few.The model implies that any set receiving more than one inputs through

a strong connection from other sets can exhibit multiple states of priming.This is so for the following reasons:

(i) An input unit Sp can reverberate at different levels.(ii) If Sq receives strong connections from set Sp, priming of Sq is, as a

consequence of different levels of reverberation, stronger or weaker

as well.

This is implied by the earlier formulation of neuronal grammar in

Chapter 10 (see also Excursus E4). According to principles (A3iii) and

(A3iv), reverberation and priming of a set leads to priming of a strongly

connected set, and, if reverberation is at levels R1, R2, and so on, the prim-ing levels are, accordingly, adjusted to P 1, P 2, and so on. Thus, levels of 

activity are, so to speak, transported through strong connections. Briefly, the

Page 260: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 260/331

12.2 Multiple Activity States of a Neuronal Set 243

following modifications of principles (A3) and (A4) are proposed:

(A3’) Activity changes caused in Sq through a connection between Sp and Sq

(i) Sp (I ,t ), Sq (0,t ) ⇒ Sq (P 1,t +1)only if Sp→→ Sq

(ii) Sp (I ,t ), Sq (R i ,t ) ⇒ Sq (I ,t +1)

only if i < 3

(iii) Sp (I ,t ), Sq (R i ,t ) ⇒ Sq (P 1 R i ,t +1)

only if Sp→→ Sq and i > 2

(iv) Sp (I ,t ), Sq (R i  R j ,t ) ⇒ Sq (I R j ,t +1)

only if Sp→→ Sq, i < 3, and j > 2

(v) Sp (R i ,t ), Sq (0,t ) ⇒ Sq (P i ,t +1)

only if Sp→→ Sq

(vi) Sp (R i ,t ), Sq (R j ,t ) ⇒ Sq (P i  R j ,t +1)

only if Sp→→ Sq, and i  = j 

(vii) Sp (R i ,t ), Sq (R i ,t ) ⇒ Sq (R i ,t +1)

only if Sp→→ Sq

(viii) Sp (P i ,t ), Sq (0,t ) ⇒ Sq (P i ,t +1)

only if Sp→→ Sq, and Sp is a sequence set

(ix) Sp (P i ,t ), Sq (P  j ,t ) ⇒ Sq (P i  P  j ,t +1)

only if Sp→→ Sq, Sp is a sequence set, and i  = j 

(x) Sp (P i ,t ), Sq (P i ,t ) ⇒ Sq (P i ,t +1)

only if Sp→→ Sq, Sp is a sequence set

(A4’) Activity changes caused in Sr through connections with Sp and Sq

(i) Sp (I ,t ), Sq (I ,t ), Sr (0,t ) ⇒ Sr (I ,t +1)

(ii) Sp (I ,t ), Sq (R i ,t ), Sr (0,t ) ⇒ Sr (I ,t +1)

only if i = 1 or i = 2, and Sq ➔ Sr

(iii) Sp (I ,t ), Sq (R i ,t ), Sr (0,t ) ⇒ Sr (P 1 P i ,t +1)

only if i > 2, Sp→→ Sr, and Sq→→ Sr

(iv) Sp (I ,t ), Sq (P i ,t ), Sr (0,t ) ⇒ Sr (I ,t +1)

only if i = 1 or i = 2, Sq→→ Sr, and Sq

is a sequence set

(v) Sp (I ,t ), Sq (P i ,t ), Sr (0,t ) ⇒ Sr (P 1 P i ,t +1)

only if i > 2, Sp→→ Sr, Sq→→ Sr, and

Sq is a sequence set(vi) Sp (P /R i ,t ), Sq (P /R j ,t ), Sr (0,t ) ⇒ Sr (P i  P  j ,t +1)

only if i > 2, Sp→→ Sr, Sq→→ Sr, (and

sets exhibiting P  at t are sequence set)

(vii) Sp (P /R i ,t ), Sq (P /R i ,t ), Sr (0,t ) ⇒ Sr (P i , t +1)

only if i > 2, Sp→→ Sr, Sq→→ Sr, (and

sets exhibiting P  at t are sequence set)

Page 261: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 261/331

244 Refining Neuronal Grammar

Briefly, according to this proposal, each neuronal set is allowed to exhibit

multiple activity states. Input units can exhibit several reverberations and

can exhibit multiple priming at a time, depending on their history of externalstimulation and the amount of priming they receive through connections in

the network. In contrast, sequence sets can exhibit ignition, reverberation,or priming, and in addition, they can exhibit multiple additional priming.

In activity tables, different activity levels of one set are written on top of 

each other, the highest activity level appearing in the top line and the lower

activity levels in the lines following (see tables in Excursuses E4 and E5).

These formulations include occasional redundancies. For example, (3iv)

follows from (3

ii) and (3

iii) if both principles are applied at the same timeand if they are assumed to act independently on different levels of reverbera-

tion. Such “additivity” of processing at different levels of activity is assumed

from now on. This amounts to considering each neuronal set as a pushdown

store with the possibility of storing several hierarchical activity tags that

are stored and retrieved independently (Pulvermuller, 1993, 2002). Only the

highest activity tags denote “visibility” of the respective activity state andhave an immediate influence on ongoing computations.

Note that in the present context activity states of sets, rather than the setsthemselves, are called visible. What has been called a visible set earlier would

now be a set with a visible activity state.

If an input unit has ignited repeatedly as a result of external input, it ex-

hibits several reverberation states at a time. If the input unit is connectedto more than one lexical category representation, more than one of these

representations can be activated – coding the information that the repeated

element in the input has been classified as member of different lexical cat-

egories. To specify inhibitory interactions between competing lexical cate-gory representations in the case of multiple reverberation of one input unit,

principle (A6) must be extended. The additional assumption proposed nextexcludes simultaneous preactivity at levels R1 or P 1 of two category repre-

sentations connected to one input unit.

(A6’) Inhibition between two lexical category representations α and β con-

nected to an active input unit S (which ignites or reverberates at R 1):

(i) Ignition of S can spread only to sets included in either α or β. The

most strongly activated representation wins.

(ii) If one or more sets included in α but not in β ignite at t , then no set

included in β and not included in α can exhibit I , P 1, or R 1 at t +1.

(iii) If all sets included in α but not in β are reverberating at R 1 at t ,thenno

set included in β and not included in α can exhibit P 1 or R 1 at t +1.

Page 262: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 262/331

12.3 Multiple Center Embedding 245

Criteria for deciding which of two alternative lexical category representa-

tions is more active (A6i) are discussed in Chapter 10. Axioms (A6ii) and

(A6

iii) would be thought to imply that activity levelsP 

1 orR

1 of sets inβ

drop to P 2 or R2, respectively. Note the modification introduced by (A6iii).

In the case of multiple reverberation of input unit S, it is now possible to havepriming effects of reverberation R1 exerted on α, whereas priming effects

of a simultaneous reverberation Ri – at a lower level of  Ri , i > 1 – can still

be effective and affect β.

Intuitively, (A6i)–(A6iii) can be attributed to a process of inhibition

between lexical category representations that may be “enabled” by an input

unit that represents an ambiguous symbol, as detailed earlier.The simulations presented in Excursuses E4 and E5 show these principles

at work on sentences.

12.3 Multiple Center Embedding

Apart from multiple occurrence of individual lexical items, the organization

of complements and adjuncts and the representation of lexical category in-

formation, a fundamental language problem that can be solved by humansbut by no other species, arises in the parsing, analyzing, and understand-

ing of center-embedded sentences. Sentence (7) is an example of how a

neuronal grammar consisting of neuronal sets can process multiple center

embeddings.

(7) Betty who John who Peter helps loves gets up.

One may question whether this is, indeed, a correct English sentence.One may hold that multiple embeddings of this kind are highly unusual,

play no role in conversations, and are considered bad style, even in written

language. Therefore, why should a grammar model focusing on the putativebrain mechanisms care about them?

The reason is historical. Chomsky (1957, 1963) rejected neuronal finite

state automata as devices for describing languages, a main reason being

that they fail at processing recursively enumerable languages with multiple

center embeddings. Context-free rewriting systems were necessary, he thenargued. This led to a paradigm shift, from neuronal (finite-state) to abstract(pushdown) automata. Sentences such as (7) were, so to speak, the flagships

of the paradigm shift.

Inthiscontext,itisclaimedthataneuronalgrammariscapableofhandling

these then crucial strings. Readers who believe that strings such as (7) are

ungrammatical may tolerate some elaboration of this issue here for historical

Page 263: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 263/331

246 Refining Neuronal Grammar

reasons. The intention is to show that this string type cannot be used to

ground arguments against neuronal models of syntax.

One further extension of the framework is necessary for allowing thenetwork to process center-embedded sentences. According to Braitenberg

(1996), some words included in the category of grammatical function wordsmay have a special neuronal representation in that their activation not only

implies excitation of a defined neuronal set, but also has an effect of the

activity state of the entire brain. An example would be words that prepare

the listener to process larger more complex sentence types. Processing of 

more complex structures may require that activity levels are lowered more

than usual to avoid the undesirable consequence of having too much activityand the concomitant interference of activity patterns (cf. Section 5.1).

Words that signal the occurrence of complex string types are, for example,

relative pronouns such as that, who, whom, or which that occur at the start

of the subordinate clause. These and other words that, so to speak, change

(or “translate”) a main sentence into a subordinate sentence have been

labeled translatives in approaches to dependency grammar (Heringer, 1996;Tesniere, 1959). Because such words warn the listener that there is much

more information to come than in an ordinary sentence, it can be postulatedthat they lead to downward regulation of brain activity. A simple way to

model this in a neuronal grammar automaton is to change the global activity

level whenever an element of the translative category occurs in the input.

Translatives such as that and who could therefore have a strong deactivat-ing effect, similar to activation of threshold control by unprimed ignition. It

is assumed here that their occurrence in the input activates threshold con-

trol regardless of their previous state of activity, and even more strongly

than unprimed full ignitions do. The translative-induced activity change islabeled I ∧∧ , implying double activation of threshold control with the ignitionin the context of translatives. Principle (A5) is adjusted to this assumption

as follows:

(A5’) Threshold regulation

(i) If a full ignition I ∧ of an input unit S happens at t , then for all rever-

berating S :

S  (R i ,t ) ⇒ S  (R i +1,t +1)(ii) If there is no ignition or reverberation R 1 at t , then for all reverberating

sets S :

S  (R i ,t ) ⇒ S  (R i +1,t +1)

(iii) External stimulation of an input unit S t representing a translative

causes double activation I ∧∧ of threshold regulation, and for all sets

S  reverberating at t :

S t (I ∧∧ ,t ) ⇒ S  (R i +2,t +1)

Page 264: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 264/331

12.4 Summary, Open Questions, and Outlook 247

Excursus E5 discusses processes in a neuronal grammar circuit induced by

a sentence with multiple center embeddings in greater detail.

12.4 Summary, Open Questions, and Outlook

This chapter refines the approach to neuronal grammar proposed in Chap-

ters 10 and 11. The main points relate to a putative neurobiological distinc-tion between obligatory complements of a word and its free adjuncts. The

adjunct relationship is proposed to be realized by unidirectional connections,

an assumption that is not in conflict with the neuroanatomical reciprocity

principle because it is based on unlearning of existing connections. Multi-ple reverberation was proposed to be possible in neuronal sets. This radical

statement is not motivated by empirical data, but the perspectives it offersappear to be a necessary component of any approach to language. There

may be alternative candidate neuronal mechanisms for this, but they must

first be found. Ignition caused by external input is proposed to lead to an

additional activity wave traveling in a word web or input unit. The machinery

that may support the processing of center-embedded sentences is discussed

in detail, although there is doubt whether this sentence type is of importancein ordinary language use. The ideas and proposals put forward in this chapter

are further explored and elaborated in the context of two more simulations

detailed in Excursuses E4 and E5.

Considering this approach to grammar in the brain as a whole, general

criticisms can be made. Its obvious limitations can be taken as desiderataand questions for future research. For example: The important question of 

how a neuronal grammar can be learned is addressed only in passing. The

postulate is that there is a small set of neuronal principles underlying syn-tactic knowledge. These principles, or at least relevant aspects of them, are

grounded in neurophysiology and neuroanatomy and are, as such, geneti-

cally determined, or at least under strong influence of the genetic code. InChapters 10–12, the postulate is that the neuroscientific principles on which

grammar rests include complex aspects of neuronal wiring and function as

follows: The specialization of neuron ensembles as sequence sets and word

webs, the distinct types of activity states of neuronal sets and their “push-

down dynamics,” the criterion for string acceptance and the other principlesformulated by principles (A1) and (A2)–(A6). Everything beyond these

neuronal mechanisms – namely, the connections between word webs and

sequence sets and the connection between different sequence sets – are as-

sumed to be the result of associative learning. At this point, it appears one

of the most important desiderata to spell out these learning mechanisms in

greater detail. Closely connected to this issue is the question of how syntacticcategory representations could develop and how their mutual inhibition is

Page 265: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 265/331

248 Refining Neuronal Grammar

established. Some preprogrammed information may be relevant here as

well. Another open question addresses the mechanisms on which the cri-

teria for string acceptance are based. What kind of neuronal device couldcheck whether the three criteria postulated are all met? These points are

relevant for developing further the present proposal.There is a multitude of important linguistic issues as well that are not

addressed here. Among them are questions about the connection between

the computation of the structure of a sentence and that of its semantics and

about mechanisms underlying the speech acts performed by using sentences.

There are open questions about dynamic relationships between word forms

and meaning, such as between pronouns and their referents, and about therelationship between the meanings of consecutive sentences. All these is-

sues are not addressed in this book. However, the general perspective on

syntactic–semantic interaction the present approach may offer should be

sketched briefly.

The proposal in Chapter 4 and elsewhere was that words are repre-

sented as distributed neuron webs. The information concerning the formand meaning of a word are processed within one functional system. Be-

cause form and meaning representations are tightly linked, the serial-ordermachinery – what is called neuronal grammar  – influences both form and

meaning representations. The input provided by sequence sets may primar-

ily reach the word form representations of word webs, but because the form

and meaning representation of a word are closely connected, the serial-order machinery ultimately exerts an influence on the semantic parts of the

word webs as well. The idea is that it is the very connection to the net-

work of sequence detectors that modifies activity in word webs in a way that

also has an effect on the semantic parts of the cell assemblies. This couldbe a neuronal route to the interaction of form-related and semantic infor-

mation. Again, this must be developed and translated into algorithms andnetworks.

From a formal linguistic point of view, networks consisting of neuronal

sets can accept many of the strings defined by one-sided linear grammars.

After all, they are a further development of finite-state networks. It has also

been argued that some sets of strings that are outside the reach of finite

state grammars (the set of center-embedded strings) and even outside thereach of context-free rewriting systems (certain sentences with subject–verb

agreement and particle verbs) can be accepted and processed successfully by

neuronal grammar. However, a more precise definition of the sets of strings

that can be processed by neuronal grammar may be desirable.

One important issue neglected here is the generation of a sentence: all

simulations focus on network dynamics induced in a network by a string in

Page 266: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 266/331

12.4 Summary, Open Questions, and Outlook 249

the input. To work out the productive side of neuronal grammar constitutes

another entry in the list of to-be-solved tasks.

As mentioned, many syntactic phenomena classic approaches to syntax

can handle are not discussed here in the context of neuronal grammar mech-

anisms. It may be useful, however, to list syntactic phenomena addressed

briefly in this chapter that have been addressed in detail on the basis of cir-

cuits. This does not imply that the solutions offered are considered optimal

or satisfactory.

   r Lexical category and subcategory   r Valence   r

Dependency between adjacent constituents   r Long-distance dependency   r Distributed word and discontinuous constituent   r Subject–verb agreement   r Adjunct relation   r Multiple word use in a sentence   r Resolution of lexical ambiguity   r Subordination   r

Multiple center embedding.

Page 267: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 267/331

EXCURSUS FOUR

Multiple Reverberation for ResolvingLexical Ambiguity

This excursus illustrates circuits motivated by the proposals discussed in

Chapter 12. The abbreviations used here and the representation of activ-

ity dynamics in table form are the same as those used in Excursuses E2

and E3.With the extensions of the grammar circuits proposed in Chapter 12, it

now becomes possible to treat sentences in which the same word occurstwice and as member of different lexical categories, such as sentence (1).

(1) Betty switches the switch on.

In Chapter 10, Figures 10.3 and 10.4 are used to sketch a putative neuronal

correlate of the syntactic category representations that may be connected

to the representation of the word form switch. Two of the lexical categories,

transitive particle verb, here abbreviated as V, and accusative noun, here

abbreviated as N, are relevant for the processing of the syntactically ambigu-ous word used in sentence (1). The homophonous words and their lexicalcategories are characterized by the assignment formulas (2) and (3) and the

valence formulas (4) and (5).

(2) switch ↔ V (transitive particle verb)

(3) switch ↔ N (accusative noun)

(4) V (p /∗/ f 1, f 2, f 3)

(5) N (p1, p2 /∗/)

Figure E4.1 shows the entire network used for sentence processing. The

representation of the ambiguous word form is doubled for ease of illus-

tration. This figure is almost identical to Figure E3.1, which dealt with a

similar sentence. Table E4.1 presents activity dynamics of the sets involved

in processing the ambiguous word and its two lexical categories. The overall

network dynamics are very similar to those in the derivation illustrated in

250

Page 268: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 268/331

Page 269: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 269/331

252 Multiple Reverberation for Resolving Lexical Ambiguity

Table E4.1. Time course of activation and deactivation of word webs andsequence sets of Figure E4.1 during processing of the string “Betty

switches the switch on,” which is congruent with the network structure.Lines represent time steps. Time steps and input to the network are listedon the left. In each column, activity states of three past-orientedsequence sets (p; to the left of slashes), of a word web (for the word form switch), and of three future-oriented sequence sets (f; to the right ofslashes) are shown. The two leftmost sequence sets (labeled Det/ and Vt/)make up the lexical category representation of an accusative noun, andthe other four sequence sets (Prop/, /Vs, /N4, /Vpart) make up that of atransitive particle verb. Time steps between inputs are chosen so that

activity spreads until stable states are reached. Letters indicate ignition(I  ), reverberation (R ), priming (P ), and rest (0). I 

∧ indicates full unprimedignition. P 1 and R 1 denote the highest levels of priming and reverberation.Numbers >1 following P or R  imply lower activity ranks, with highernumbers indicating lower activity levels.

N(p1) N(p2) V(p) V(f  1) V(f  2) V(f  3)

Det/ Vt/ Prop/ /switch/ /Vs /N /Vpart

0 0/ 0/ 0/ /0/ /0 /0 /0

1 Betty input unit of Pn “Betty” ignites (full I∧)2 0/ 0/ P1/ /0/ /0 /0 /0

3 0/ 0/ P1/ /P1/ /0 /0 /0

4 switch 0/ 0/ P1/ /I/ /0 /0 /0

5 0/ 0/ I/ /R1/ /P1 /P1 /P1

6 0/ P1 / R1/ /R1/ /P1 /P1 /P1

7 -es input unit of Vs “-es” ignites (primed I)

8 0/ P1/ R1/ /R1/ /I /P1 /P1

9 0/ P1/ R1/ /I/ /R1 /P1 /P1

10 0/ P1/ I/ /R1/ /R1 /P1 /P1

11 the input unit of “the” ignites (full I∧)

12 P1/ P2/ R2/ /R2/ /R2 /P2 /P2

13 P1/ P2/ R2/ /P1 / /R2 /P2 /P2

 /R2 /

14 switch P1/ P2/ R2/ /I/ /R2 /P2 /P2

 /R2 /

15 I/ I/ R2/ /R1/ /R2 /P2 /P2

/R2/

16 R1/ R1/ R2/ /R1/ /R2 /I /P2

/R2/

17 R2/ R2/ R2/ /I/ /R2 /R1 /P2

/R2/

18 R2/ R2/ I/ /R1/ /I /R1 /P1

/R2/ /P2 /P2

Page 270: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 270/331

Multiple Reverberation for Resolving Lexical Ambiguity 253

N(p1) N(p2) V(p) V(f  1) V(f  2) V(f  3)

Det/ Vt/ Prop/ /switch/ /Vs /N /Vpart

19 R2/ P1/ R1/ /R1/ /R1 /R1 /P1

R2/ /R2/ /P2 /P2 /P2

20 on input unit of “on” ignites (primed I)

21 R2/ R2/ R1/ /R1/ /R1 /R1 /I

/R2/ /P2 /P2 /P2

22 R2/ R2/ R1/ /I/ /R1 /R1 /R1

/R2/ /P2 /P2 /P2

23 R2/ R2/ I/ /R1/ /I /I /R1/R2/ /P2 /P2

24 R2/ I/ R1/ /R1/ /R1 /R1 /R1

/R2/ /P2 /P2 /P2

25 R1/ R1/ R2/ /I/ /R2 /R2 /R2

P2/ /R2/

26 I/ R1/ R2/ /R1/ /R2 /R2 /R2

P2/ /R2/

27 R1/ R1/ R2/ /R1/ /R2 /R2 /R2P2/ /R2/

28 DEACTIVATE 0/ 0/ 0/ /0/ /0 /0 /0

Two of the sequence sets directly connected with the input unit of the

ambiguous word form switch are also directly connected to each other. The

backward-looking sequence detector of the N designed to detect a preceding

V are directly connected to the forward-looking sequence detector of V

sensitive to a follower N on the right. This leads to a circle in the network,as sketched by (6).

(6) switch →→ V (f 2) →→ N (p2) →→ switch

The circle can give rise to reverberation of activity between neuronal sets.

It is this loop that explains why, after activating the verb representation

together with the word form switch, the network later tends to classify it

as a noun. At time step 6, N (p2) is being primed through this loop, and attime steps 14–15 and 23–24, ignitions spread through it, notably in different

directions.

The derivation from time steps 1–5 is standard. The first occurrence of 

the word form is classified as a verb, because it has been preceded by a

proper name (Prop). The verb suffix (Vs) provides further support for this

classification (steps 8–10). The determiner (Det) in the input causes full

Page 271: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 271/331

254 Multiple Reverberation for Resolving Lexical Ambiguity

ignition and therefore reduction of all activity levels. At this point, sequence

sets being part of the lexical category representation of N have accumulated

more activity [they exhibit priming levels of P1 and P2 (average: 1.5), respec-tively, whereas all activity levels of the competing category are at P2 or R2

(average: 2)].Therefore, the next ignition of the input unit representing the word form

 switch ignites only the sets included in the N representation [according to

(A6i) – see time step 14]. Notably, classification of the second occurrence

of the repeated word as an N provides further support for the correctness

of the classification of its first occurrence as a verb. This is expressed by the

ignition of V (f 2) at step 16, which leads to ignitions of almost all sequencesets included in the verb category representation. This ignition also reduces

the activity levels of the competing and reverberating lexical category rep-

resentation [as stated by (A6iii)].

After presentation of the final word in the sentence, a wave of ignitions

spreads through the network causing synchronization of the already visible

and satisfied sets. It is noteworthy that during the spreading process, the hier-archy of reverberation of the two competing lexical category representations

is again being changed (time steps 24–25).This example demonstrates that a network composed of neuronal sets

can process sentences in which a word form occurs twice but as a member

of different lexical categories. The processing not only includes “accepting”

the string, but a spatiotemporal pattern of activity also arises that includesinformation about the hierarchical organization of the sentence and the

assignment of morpheme tokens to lexical categories. Most notable is the

dynamic linking of the word-form representation with its alternative lexical

category representations in the course of the processing of a sentence thatincludes an ambiguous word.One may still ask many further questions, however; for example, about

how three or more occurrences of the same word form would be processed.

Nevertheless, the proposed framework poses no a priori limit to the number

of occurrences of a word, and there appears to be no principled reason

why a network composed of neuronal sets should be unable to deal with

a construction in which a given word form occurs more than twice. The

introduced examples may suffice for illustrating perspectives of the presentapproach.

Page 272: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 272/331

Page 273: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 273/331

256 Multiple Reverberations and Multiple Center Embeddings

Figure E5.1. Fragment of neuronal grammar that processes and accepts the sentence“Betty who John who Peter helps loves gets up.” (An equivalent formalism is providedin the text.) Activation and deactivation dynamics of the network are shown in Table E5.1.

(3) Betty, John, Peter ↔ Prop (nominative proper name)

(4) helps, loves ↔ Vsub (subordinate transitive verb)

(5) gets ↔ Vip (intransitive particle verb)

(6) up ↔ Vpart (verb particle)

With these assignment formulas, a simplification is made compared with

the grammars in earlier chapters. Verb suffixes are again treated as parts

of the verbs. This is done to shorten the derivation. Previous examples had

shown how noun–verb agreement is modeled. The same applies to a more

elaborate version of the present grammar.

The lexical categories are characterized by the following valence

properties:

(7) Tr (/∗/ f)

(8) Prop (/∗/ f)

(9) Vsub (p1, p2 /∗/)

(10) Vip (p /∗/ f)

(11) Vpart (p /∗/)

Page 274: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 274/331

Multiple Reverberations and Multiple Center Embeddings 257

This proposal implies that verbs in main clauses and in subordinate sentences

are attributed to distinct lexical categories. The main clause verb would

search for its complements in the past and future, so to speak, whereas averb in a subordinate clause would restrict search to the past.

Connections between sequence sets are expressed by the following se-quence formulas:

(12) Tr (f) →→  Vsub (p2)

(13) Prop (f)→→  Vsub (p1)

(14) Prop (f)→→  Vip (p)

(15) Vip (f) →→  Vpart (p)

Formulas (13) and (14) denote the same connection; sets Vsub (p1) and Vip

(p) are considered identical. Therefore, these formulas could, in principle,

be replaced by (16), which then refers to a connection shared between thetwo alternative representations of the two distinct lexical category represen-tations of verbs appearing in main and subordinate sentences, respectively.

Also, the sequence detector V(p) is shared by all verb representations.

(16) Prop (f)→→  V (p)

Note again that in the present grammar fragment, a distinction is made

between ordinary verbs in main clauses and verbs in subordinate sentences

(Vsub), which are assumed to take translativesas their complements. Transla-

tives take the place of the accusative noun as the verb complement. However,in contrast to ordinary accusative nouns that must appear to the right of the

verb, translatives are placed to the left of transitive verbs. This motivates the

postulate of a distinct set of sequence sets for the new lexical category Vsub.To model this, some grammars introduce transformations, movements, or

other sophisticated and costly procedures. In the proposed neuronal gram-

mars, the solution is different. Each transitive verb is supplied with twodistinct sets of sequence features. The “standard” form (10) is characterized

by nominative and accusative complements occurring to the left and right

of the verb, respectively, and the probably less-frequent subordinate form

(9), in which both complements occur to the left and one complement is a

translative.Postulating a small number of additional sets of sequence features for

lexical categories appears less costly than introducing new operations, such

as transformations or movements (see Chapter 10, Section 10.6.4). After

all, the number of sentence forms consisting of the same lexical categories

but differing in their order is small in English. Introduction of distinct sets

of sequence features for verbs occurring in sentence forms such as active,

Page 275: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 275/331

       T     a       b       l     e

       E       5  .       1  .     T     i    m    e    c    o    u    r    s    e    o      f    a    c    t     i    v    a    t     i    o    n    a    n     d     d    e    a    c    t     i    v    a    t     i    o    n    o      f    w    o    r     d    w    e     b    s    a    n     d    s    e    q    u    e    n    c    e    s    e    t    s    o      f     F     i    g    u    r    e     E     5 .     1

     d

    u    r     i    n    g

    p    r    o    c    e    s

    s     i    n    g    o      f    t     h    e    s    t    r     i    n    g     “     B    e    t    t    y

    w     h    o     J    o     h    n    w     h    o     P    e    t    e    r     h    e

     l    p    s     l    o    v    e    s    g    e    t    s    u    p ,     ”    w     h     i    c     h

     i    s    c    o    n    g    r    u    e    n    t    w     i    t     h    t     h    e

    n    e    t    w    o

    r     k    s    t    r    u    c    t    u    r    e .

     L     i    n    e    s    r    e    p    r    e    s

    e    n    t    t     i    m    e    s    t    e    p    s .

     T     i    m    e    s    t    e

    p    s    a    n     d     i    n    p    u    t    t    o    t     h    e    n    e    t    w

    o    r     k    a    r    e     l     i    s    t    e     d    o    n    t     h    e     l    e      f    t

 .     I    n

    e    a    c     h    c

    o     l    u    m    n ,    a    c    t     i    v     i    t    y    s    t    a    t    e    s    o      f

    o    n    e    o    r    m    o    r    e    o    p    t     i    o    n    a     l    p    a

    s    t  -    o    r     i    e    n    t    e     d    s    e    q    u    e    n    c    e    s    e    t    s     (    p    ;    t    o    t     h    e     l    e      f    t    o      f    s     l    a    s     h    e

    s     ) ,    o      f

    a    w    o    r     d

    w    e     b     (     b    e    t    w    e    e    n    s     l    a    s     h    e    s     ) ,

    a    n     d    o      f    o    p    t     i    o    n    a     l      f    u    t    u    r    e  -    o

    r     i    e    n    t    e     d    s    e    q    u    e    n    c    e    s    e    t    s     (      f    ;    t    o    t     h    e    r     i    g     h    t    o      f    s     l    a    s     h    e    s     )    a

    r    e

    s     h    o    w    n

 .     T     i    m    e    s    t    e    p    s     b    e    t    w    e    e    n     i    n    p    u    t    s    a    r    e    c     h    o    s    e    n    s    o    t     h    a    t    a

    c    t     i    v     i    t    y    s    p    r    e    a     d    s    u    n    t     i     l    s    t    a     b

     l    e    s    t    a    t    e    s    a    r    e    r    e    a    c     h    e     d .

     L    e

    t    t    e    r    s

     i    n     d     i    c    a    t    e     i    g    n     i    t     i    o    n     (       I     ) ,    r    e    v    e    r     b    e    r    a    t     i    o    n     (       R     ) ,    p    r     i    m     i    n    g     (       P     ) ,    a    n     d    r    e    s    t     (     0     ) .       I     ∧

     i    n     d     i    c    a    t    e    s      f    u     l     l    u

    n    p    r     i    m    e     d     i    g    n     i    t     i    o    n ,    a    n     d       I     ∧

     ∧

    a    n

     i    g    n     i    t     i    o

    n    c    a    u    s     i    n    g    s    t    r    o    n    g    t     h    r    e    s     h    o

     l     d    r    e    g    u     l    a    t     i    o    n .

       P     1

    a    n     d       R     1

     d    e    n    o    t    e    t     h    e     h     i    g     h    e    s    t    a    c    t     i    v     i    t    y     l    e    v    e     l .     N    u    m     b    e    r    s    >     1

      f    o     l     l    o    w     i    n    g       P

    o    r       R

    m    e    a    n    a    c    t     i    v     i    t    y    r    a    n     k    s ,    w     i    t     h     h     i    g     h    e    r    n    u    m     b    e

    r    s     i    n     d     i    c    a    t     i    n    g     l    o    w    e    r     l    e    v    e     l    s .

    /    W    h   o    /    f

    /    B   e    t    t   y    /    f

    /    J   o    h   n    /    f

    /    P   e    t   e   r    /    f

   p    1 ,   p    2    /    H   e    l   p   s    /

   p    1 ,   p    2

    /    L   o   v   e   s    /

   p    /    G   e    t   s    /    f

    l    /    U   p    /

      0

      /      0      /      0

      /      0      /      0

      /      0      /      0

      /      0      /      0

      0 ,      0

      /      0      /

      0 ,      0

      /      0

      /

      0      /      0      /      0

      0      /      0      /

      1      B    e     t     t    y

      /      0      /      0

      /      I     ∧      /      0

      /      0      /      0

      /      0      /      0

      0 ,      0

      /      0      /

      0 ,      0

      /      0

      /

      0      /      0      /      0

      0      /      0      /

      2

      /      0      /      0

      /      R      1      /      P      1

      /      0      /      P      1

      /      0      /      P      1

      0 ,      0

      /      0      /

      0 ,      0

      /      0

      /

      0      /      0      /      0

      0      /      0      /

      3

      /      0      /      0

      /      R      1      /      P      1

      /      0      /      P      1

      /      0      /      P      1

      P      1 ,      0

      /      0      /

      P      1 ,      0      /

      0      /

      P      1      /      0      /      0

      0      /      0      /

      4

      /      0      /      0

      /      R      1      /      P      1

      /      0      /      P      1

      /      0      /      P      1

      P      1 ,      0

      /      0      /

      P      1 ,      0      /

      0      /

      P      1      /      P      1      /      0

      0      /      0      /

      5    w      h    o

      /      I     ∧     ∧      /      0

      /      R      3      /      P      3

      /      0      /      P      3

      /      0      /      P      3

      P      3 ,      0

      /      0      /

      P      3 ,      0      /

      0      /

      P      3      /      P      3      /      0

      0      /      0      /

     6

      /      R      1      /      P      1

      /      R      3      /      P      3

      /      0      /      P      3

      /      0      /      P      3

      P      3 ,      0

      /      0      /

      P      3 ,      0      /

      0      /

      P      3      /      P      3      /      0

      0      /      0      /

      7

      /      R      1      /      P      1

      /      R      3      /      P      3

      /      0      /      P      3

      /      0      /      P      3

      P      3 ,      P      1      /      0      /

      P      3 ,      P      1

      /      0      /

      P      3      /      P      3      /      0

      0      /      0      /

      8

      /      R      1      /      P      1

      /      R      3      /      P      3

      /      0      /      P      3

      /      0      /      P      3

      P      3 ,      P      1      /      P      2      /

      P      3 ,      P      1

      /      P      2      /

      P      3      /      P      3      /      0

      0      /      0      /

      9      J    o      h    n

      /      R      2      /      P      2

      /      R      4      /      P      4

      /      I     ∧

      /      P      4

      /      0      /      P      4

      P      4 ,      P      2      /      P      3      /

      P      4 ,      P      2

      /      P      3      /

      P      4      /      P      4      /      0

      0      /      0      /

      1      0

      /      R      2      /      P      2

      /      R      4      /      P      1

      /      R      1      /      P      1

      /      0      /      P      1

      P      4 ,      P      2      /      P      3      /

      P      4 ,      P      2

      /      P      3      /

      P      4      /      P      4      /      0

      0      /      0      /

      /      P      4

      /      P      4

      /      P      4

      1      1

      /      R      2      /      P      2

      /      R      4      /      P      1

      /      R      1      /      P      1

      /      0      /      P      1

      P      1 ,      P      2      /      P      3      /

      P      1 ,      P      2

      /      P      3      /

      P      1      /      P      4      /      0

      0      /      0      /

      /      P      4

      /      P      4

      /      P      4

      P      4 ,

      /

      P      4 ,      /

      P      4      /

258

Page 276: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 276/331

      1      2

      /      R      2      /      P      2

      /      R      4      /      P      1

      /      R      1      /      P      1

      /      0      /      P      1

      P      1 ,      P      2      /      P      1      /

      P      1 ,      P      2

      /      P      1      /

      P      1      /      P      1      /      0

      0      /      0      /

      /      P      4

      /      P      4

      /      P      4

      P      4 ,

      /

      P      4 ,      /

      P      4      /

      1      3    w      h    o

      /      I     ∧     ∧      /      P      4

      /      R     6

      /      P      3

      /      R      3      /      P      3

      /      0      /      P      3

      P      3 ,      P      4      /      P      3      /

      P      3 ,      P      4

      /      P      3      /

      P      3      /      P      3      /      0

      0      /      0      /

      /      R      4      /

      /      P     6

      /      P     6

      /      P     6

      P     6 ,

      /

      P     6 ,      /

      P     6

      /      P     6

      /

      1      4

      /      R      1      /      P      1

      /      R     6

      /      P      3

      /      R      3      /      P      3

      /      0      /      P      3

      P      3 ,      P      4      /      P      3      /

      P      3 ,      P      4

      /      P      3      /

      P      3      /      P      3      /      0

      0      /      0      /

      /      R      4      /      P      4

      /      P     6

      /      P     6

      /      P     6

      P     6 ,

      /

      P     6 ,      /

      P     6

      /      P     6

      /

      1      5

      /      R      1      /      P      1

      /      R     6

      /      P      3

      /      R      3      /      P      3

      /      0      /      P      3

      P      3 ,      P      1      /      P      3      /

      P      3 ,      P      1

      /      P      3      /

      P      3      /      P      3      /      0

      0      /      0      /

      /      R      4      /      P      4

      /      P     6

      /      P     6

      /      P     6

      P     6 ,      P      4      /

      P     6 ,      P      4

      /

      P     6

      /      P     6

      /

      1     6

      /      R      1      /      P      1

      /      R     6

      /      P      3

      /      R      3      /      P      3

      /      0      /      P      3

      P      3 ,      P      1      /      P      2      /

      P      3 ,      P      1

      /      P      2      /

      P      3      /      P      3      /      0

      0      /      0      /

      /      R      4      /      P      4

      /      P     6

      /      P     6

      /      P     6

      P     6 ,      P      4      /      P      5      /

      P     6 ,      P      4

      /      P      5

      P     6

      /      P     6

      /

      1      7      P    e     t    e    r

      /      R      2      /      P      2

      /      R      7      /      P      4

      /      R      4      /      P      4

      /      I     ∧

      /      P      4

      P      4 ,      P      2      /      P      3      /

      P      4 ,      P      2

      /      P      3      /

      P      4      /      P      4      /      0

      0      /      0      /

      /      R      4      /      P      5

      /      P      7

      /      P      7

      /      P      7

      P      7 ,      P      5      /      P     6

      /

      P      7 ,      P      5

      /      P     6

      P      7      /      P      7      /

      1      8

      /      R      2      /      P      2

      /      R      7      /      P      1

      /      R      4      /      P      1

      /      R      1      /      P      1

      P      4 ,      P      2      /      P      3      /

      P      4 ,      P      2

      /      P      3      /

      P      4      /      P      4      /      0

      0      /      0      /

      /      R      5      /      P      5

      /      P      4

      /      P      4

      /      P      4

      P      7 ,      P      5      /      P     6

      /

      P      7 ,      P      5

      /      P     6

      /

      P      7      /      P      7      /

      /      P      7

      /      P      7

      /      P      7

      1      9

      /      R      2      /      P      2

      /      R      7      /      P      1

      /      R      4      /      P      1

      /      R      1      /      P      1

      P      1 ,      P      2      /      P      3      /

      P      1 ,      P      2

      /      P      3      /

      P      1      /      P      4      /      0

      0      /      0      /

      /      R      5      /      P      5      /

      /      P      4

      /      P      4

      /      P      4

      P      4 ,      P      5      /      P     6

      P      4 ,      P      5

      /      P     6

      /

      P      4      /      P      7      /

      /      P      7

      /      P      7

      /      P      7

      P      7 ,

      /

      P      7 ,      /

      P      7      /

      2      0

      /      R      2      /      P      2

      /      R      7      /      P      1

      /      R      4      /      P      1

      /      R      1      /      P      1

      P      1 ,      P      2      /      P      1      /

      P      1 ,      P      2

      /      P      1      /

      P      1      /      P      1      /      0

      0      /      0      /

      /      R      5      /      P      5

      /      P      4

      /      P      4

      /      P      4

      P      4 ,      P      5      /      P      4

      P      4 ,      P      5

      /      P      4      /

      P      4      /      P      4      /

      /      P      7

      /      P      7

      /      P      7

      P      7 ,

      /

      P      7 ,      /

      P      7      /      P      7      /

      2      1      h    e      l    p    s

      /      R      2      /      P      2

      /      R      7      /      P      1

      /      R      4      /      P      1

      /      R      1      /      P      1

      P      1 ,      P      2      /      I      /

      P      1 ,      P      2

      /      P      1      /

      P      1      /      P      1      /      0

      0      /      0      /

      /      R      5      /      P      5

      /      P      4

      /      P      4

      /      P      4

      P      4 ,      P      5      /      P      4      /

      P      4 ,      P      5

      /      P      4      /

      P      4      /      P      4      /

      /      P      7

      /      P      7

      /      P      7

      P      7 ,

      /

      P      7 ,      /

      P      7      /      P      7      /

(     c     o     n      t       i     n

     u     e       d

      )

259

Page 277: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 277/331

       T     a       b       l     e

       E       5  .       1

                 (     c     o     n

      t       i     n     u     e       d                 )

    /    W    h   o    /    f

    /    B   e    t    t   y    /    f

    /    J   o    h   n    /    f

    /    P   e    t   e   r    /    f

   p    1 ,   p    2    /    H   e    l   p   s    /

   p    1 ,   p    2    /    L   o   v   e   s    /

   p    /    G   e    t   s    /    f

    l    /    U   p    /

      2      2

      /      R

      2      /      P

      2

      /      R

      7      /      P

      1

      /      R

      4      /      P

      1

      /      R

      1      /      P

      1

      I ,      I      /      R

      1      /

      I ,      I      /      P

      1      /

      I      /      P

      1      /      0

      0      /      0      /

      /      R

      5      /      P

      5

      /      P

      4

      /      P

      4

      /      P

      4

      P      4

 ,      P      5

      /      P

      4      /

      P      4

 ,      P      5

      /      P

      4      /

      P      4

      /      P

      4      /

      /      P

      7

      /      P

      7

      /      P

      7

      P      7

 ,      /

      P      7

 ,      /

      P      7

      /      P

      7      /

      2      3

      /      R

      2      /      I

      /      R

      7      /      I

      /      R

      4      /      I

      /      R

      1      /      I

      R      1 ,      R

      1      /      R

      1      /

      R      1 ,      R

      1      /      P

      1      /

      R      1

      /      P

      1      /      0

      0      /      0      /

      /      R

      5      /      P

      5

      /      P

      4

      /      P

      4

      /      P

      4

      P      4 ,      P

      5      /      P

      4      /

      P      4 ,      P

      5      /      P

      4      /

      P      4

      /      P

      4      /

      /      P

      7

      /      P

      7

      /      P

      7

      P      7

 ,      /

      P      7

 ,      /

      P      7

      /      P

      7      /

      2      4

      /      I      /      R

      1

      /      R

      7      /      R

      1

      /      R

      4      /      R

      1

      /      I      /      R

      1

      R      1

 ,      R      1

      /      R

      1      /

      R      1

 ,      R      1

      /      P

      1      /

      R      1

      /      P

      1      /      0

      0      /      0      /

      /      R

      5      /      P

      5

      /      P

      4

      /      P

      4

      /      P

      4

      P      4

 ,      P      5

      /      P

      4      /

      P      4

 ,      P      5

      /      P

      4      /

      P      4

      /      P

      4      /

      /      P

      7

      /      P

      7

      /      P

      7

      P      7 ,

      /

      P      7 ,      /

      P      7

      /      P

      7      /

      2      5

    /    R      1    /    R

      1

      /      R

      7      /      R

      1

      /      R

      4      /      R

      1

    /    R      1    /    R

      1

    R      1 ,    R

      1    /    R

      1    /

      R      1 ,      R

      1      /      P

      1      /

      R      1

      /      P

      1      /      0

      0      /      0      /

      /      R

      5      /      P

      5

      /      P

      4

      /      P

      4

      P      4 ,      P

      5      /      P

      4      /

      P      4 ,      P

      5      /      P

      4      /

      P      4

      /      P

      4      /

      /      P

      7

      /      P

      7

      /      P

      7

      P      7

 ,      /

      P      7

 ,      /

      P      7

      /      P

      7      /

      2      6      D      E      A      C      T      I      V      A      T      E

      /      R

      5      /      P

      5

      /      R

      7      /      P

      4

      /      R

      4      /      P

      4

      /      0      /      P

      4

      P      4

 ,      P      5

      /      P

      4      /

      P      4

 ,      P      5

      /      P

      4      /

      P      4

      /      P

      4      /      0

      0      /      0      /

      /      P

      7

      /      P

      7

      /      P

      7

      P      7

 ,      /

      P      7

 ,      /

      P      7

      /      P

      7      /

      2      7      L      E      V      E      L      A      D      J      U      S

      T

      /      R

      2      /      P

      2

      /      R

      4      /      P

      1

      /      R

      1      /      R

      1

      /      0      /      R

      1

      P      1 ,      P

      2      /      P

      1      /

      P      1 ,      P

      2      /      P

      1      /

      P      1

      /      P

      1      /      0

      0      /      0      /

      /      P

      4

      /      P

      4

      /      P

      4

      P      4 ,

      /

      P      4 ,      /

      P      4

      /      P

      4      /

      2      8      l    o    v    e    s

      /      R

      2      /      P

      2

      /      R

      4      /      R

      1

      /      R

      1      /      R

      1

      /      0      /      R

      1

      P      1

 ,      P      2

      /      P

      1      /

      P      1

 ,      P      2

      /      I      /

      P      1

      /      P

      1      /      0

      0      /      0      /

      /      P

      4

      /      P

      4

      /      P

      4

      P      4

 ,      /

      P      4

 ,      /

      P      4

      /      P

      4      /

      2      9

      /      R

      2      /      P

      2

      /      R

      4      /      R

      1

      /      R

      1      /      R

      1

      /      0      /      R

      1

      I ,      I      /      P

      1      /

      I ,      I      /      R

      1      /

      I      /      P

      1      /      0

      0      /      0      /

      /      P

      4

      /      P

      4

      /      P

      4

      P      4 ,

      /

      P      4 ,      /

      P      4

      /      P

      4      /

      3      0

      /      R

      2      /      I

      /      R

      4      /      I

      /      R

      1      /      I

      /      0      /      I

      R      1 ,      R

      1      /      P

      1      /

      R      1 ,      R

      1      /      R

      1      /

      R      1

      /      P

      1      /      0

      0      /      0      /

      /      P

      4

      /      P

      4

      /      P

      4

      P      4 ,

      /

      P      4 ,      /

      P      4

      /      P

      4      /

      3      1

      /      I      /      R

      1

      /      R

      4      /      R

      1

      /      I      /      R

      1

      /      0      /      R

      1

      R      1

 ,      R      1

      /      P

      1      /

      R      1

 ,      R      1

      /      R

      1      /

      R      1

      /      P

      1      /      0

      0      /      0      /

      /      P

      4

      /      P

      4

      /      P

      4

      P      4

 ,      /

      P      4

 ,      /

      P      4

      /      P

      4      /

260

Page 278: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 278/331

Page 279: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 279/331

262 Multiple Reverberations and Multiple Center Embeddings

passive, question, and relative clause therefore do not lead to too heavy a

burden on the grammar machinery. It is also important to recall that intro-

ducing new sets of sequence sets is not a problem in the present framework.The development and definition of a system of lexical categories along these

lines is left for future research.The possible brain basis of free adjuncts of a sentence constituent is ad-

dressed in Section 12.1. In the present example sentence, the translatives

are adjuncts that can be freely added to any noun. Thus, their respective

neuronal sets should not exhibit a strong reciprocal connection with the

neuronal set representing the word whose adjunct they can be considered.

The sets realizing noun and transitive adjunct would connect via weak unidi-rectional links. These allow for contributing to the reactivation of the mother

noun after a subordinated sentence has been processed. The present simu-

lation does not cover adjunct connections because their introduction ne-

cessitates further elaboration of the algorithm and table. Because in the

present formulation, reactivation of the mother is possible on the basis

of threshold regulation [axiom (AS)], such elaboration appears avoidablehere.

Example sentence (1) exhibits a feature discussed in Section 12.2, the mul-tiple occurrence of a word form. In this case, the translative is used twice in

the course of the analysis of a sentence. In the example simulation shown in

Table E5.1, its double use is reflected in the double activation of its neuronal

set starting at time step 13 and ending at time step 25. Because in this sim-ulation the pushdown mechanism separates the processing of the different

levels of embedding of the sentence, one of the transitive occurrences – or

tokens is processed together with the first embedded sentence and the other

occurrence with the innermost sentence embedded in both others. This isanother illustration of how multiple occurrences of the same item can beprocessed in a neuronal grammar. Note the dynamical grouping of active

representations into sets of sets that interact.

Further crucial steps in the derivation displayed in Table E5.1, the net-

work’s processing of a sentence with multiple center embeddings, are as

follows: Ignitions of input units of translatives occur at time steps 5 and

13. As a consequence, there is double activation of threshold control, and

therefore all “visible” activity states of other sets in the network become“invisible” – that is, their activity levels are reduced by two levels. At time

step 5, reduction is to levels R3 and P 3, and at time step 13, even levels be-

comeaslowas R6 and P 6. In other words, the information about the elements

that were most active in the device’s memory is now being “pushed down”

to a lower level of activity, where the activity states are no longer “visible.”Once again, this is analogous to the operation of a pushdown mechanism.

Page 280: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 280/331

Multiple Reverberations and Multiple Center Embeddings 263

According to the modified proposal in Chapter 12, each neuronal set can be

considered to include independent pushdown memories.

Because ignitions can be transmitted only through the network to visiblesets, or, more precisely, to visible activity levels of sets, ignitions can only

affect the sets reverberating and primed at the highest levels of  R1, P 1,R2, and P 2 (see Chapter 10 for discussion). After all the input units of 

one of the three embedded sentences have been activated, the respective

neuronal elements whose states are visible are synchronized and satisfied.

This becomes effective separately for the three sentences embedded into

each other, at time steps 25, 32, and 46, respectively. In the table, the relevant

activity levels at these time steps are highlighted in bold print. After separatesynchronization of each of the three sentences, the respective synchronized

parts of the complex sentence, that is, the embedded clauses, are deactivated

sequentially (time steps 26, 33, and 47). The three sentences are therefore

processed separately, as if they were not embedded into each other, but

appeared in an ordinary manner one after the other instead. Note, however,

that this analysis leaves open the question of coreference relations betweennouns and translatives.

After deactivation of the representations of embedded sentences, sev-eral sets reverberate or are primed at low levels of R or P . At this point,

the threshold machinery (A5) must increase activity levels so that the most

active sets become visible. As already mentioned, however, strictly speak-

ing, this adjustment of levels is redundant because according to the presentassumptions the highest activity levels are visible, regardless of what their

absolute values actually are (cf. Chapter 10). If the most strongly activated

sets are deactivated, the sets with the next lower levels of activity automa-

tically become visible. Thus, the changes at time steps 27 and 34 are simplythe result of relabeling of activity levels rather than the result of real activitychanges. Nevertheless, as emphasized in Section 5.1, from a neurobiological

perspective it appears necessary to postulate a mechanism for the regula-

tion of activity levels of neuronal elements by both decreasing and increasing

activity. Also, this appears advantageous for yielding activity values within

the range of optimal functioning. For these reasons, the assumption of a

threshold regulation device that can provide both global activation and de-

activation is maintained, although one aspect of it may be considered redun-dant in the present derivations. In real neuronal structures, the regulation

device may have the effect that the visible activity levels are always at about

the same level.

Several problems cannot be solved by the present network and exten-

sions are necessary. As earlier examples of grammar circuits, the presentgrammar fragment again produces some false-positive errors. This may be

Page 281: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 281/331

264 Multiple Reverberations and Multiple Center Embeddings

improved by more fine-grained adjustments. The transitives would, accord-

ing to the present concept, provide the main mechanism for separating the

processing of the embedded constructions. However, there is the problem of disambiguation of translatives, because, for example, the word formwho can

be used not only as translative connecting a superordinate to a subordinatesentence but also as an interrogative pronoun in who questions. This may

not appear to be a difficult problem, and the neuronal connections modeling

the adjunct relation and the priming they support may offer perspectives on

solving it. Yet an even worse problem appears to be that overt translatives

are not obligatory in English, as in string (17).

(17) The woman John loves gets up.

For these cases, a mechanism must be specified that does not rely on overt

translatives. One possibility is to postulate that in this case, sequence de-

tectors strongly activate threshold regulation instead of individual words. Afurther open question concerns the interaction of the proposed serial-order

machinery with the semantic representations of words. These and numerous

other aspects must be addressed in the future (Crespi-Reghizzi et al., 2001).

After considering this example, the following general concluding remarkappears appropriate. Neuronal grammar can accept strings, including center

embeddings, by grouping embedded phrases into synchronized assemblies of neuronal sets. This provides a basic structural analysis of complex sentences.

It is important that there is, in the present grammar, no upper limit for

the number of embeddings or the number of elements assigned to any given

lexical category whose representation is being activated multiply, in principle

without upper limit for the number of reverberations. Thus, there is the

possibility of producing longer and longer strings of lower and lower levelsof embedding, and new lexical elements can easily be added to the grammar

fragment. There is no theoretical reason for excluding such excessive use of 

the pushdown machinery. The only reasons are empirical in nature.

From this excursus, it appears relevant to keep in mind that neuronal

grammar can accept at least certain subsets of strings defined by context-free rewriting systems. It appears that the flagship sentence structure from

the linguistic wars against neuron-based finite state grammars can travel

calmly in the light of neuroscientific knowledge and theorizing.

Page 282: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 282/331

CHAPTER THIRTEEN

Neurophysiology of Syntax 

Models are only as good as the tools available for testing them. Is there

any perspective on testing the proposals about syntactic brain processes

discussed to this point?

13.1 Making Predictions

One obvious prediction of the neuronal grammar framework is that the

cortical processes following a sentence in the input are of different types.

Such process types include the following:

   r Ignitions of word webs   r Ignitions of sequence sets   r Downward or upward regulation of activity by threshold control   r Reverberation and priming of neuronal sets

It may be possible to investigate these putative processes using neurophys-

iological recording techniques. Because the relevant activation and deact-

ivation processes likely occur within tenths or hundredths of a second, it is

advisable to use fast neuroimaging techniques in the investigation of serial-

order information processing.

At first glance, one straightforward way of transforming any of the men-

tioned simulations into predictions about neurophysiological processes ap-

pears to be by calculating sum activity values at any time step of a simulation,and by thereby obtaining a series of activity values.

However, this strategy is not so straightforward as it appears at first glance

because several parameters must be chosen beforehand. The relative weight-

ing of the activity states is one set of parameters that appears to distinguish

between pronounced activity changes – those induced by ignitions and reg-

ulation processes – and smaller ones – as, for example, those caused by the

265

Page 283: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 283/331

266 Neurophysiology of Syntax

spreading of priming or reverberation. As an additional set of parameters,

the time constants of individual processes must be specified. The time con-

stants of the postulated processes, such as an ignition and the spreading of priming, must be chosen. Furthermore, in Excursuses E4 and E5, full ignit-

ions and the subsequent regulation of cortical activity contracted into onetime step. This is reasonable to avoid lengthy simulations, but because the

regulation process is the consequence of the strong activation process, it is

necessary to separate the two processes within a model designed to make

exact predictions on the time course of cortical activation and deactivation.

Predictions from the simulations cannot be 1 : 1 conversions.

13.2 Neuronal Grammar and Syntax-Related Brain Potentials

A principled prediction made by the model is that after a word is present in

the input, the word web ignites before its connected sequence sets. However,

this prediction is not easy to test. Because preactivated sequence detectors

can give rise to the priming of word webs, the same word appearing in

different contexts not only activates different sets of sequence sets, but the

ignition of its own word web may also differ as a function of context. It isof utmost importance here to choose appropriate baseline conditions for

comparison.

A further clear prediction is that syntactically well-formed and ungram-

matical wordstrings elicitdifferential brain responses. The processes induced

in circuits of neuronal grammar by syntactically well-formed (congruent)sentences can be compared with the effect of syntactic violations in (incon-

gruent) strings. As a first step, this can be done purely on the basis of the

model simulations. Comparison of these processes (see Excursus E2) showsthat the physiological differences between grammatical and ungrammatical

strings should be three-fold. A situation is considered in which one particular

word or morpheme fits in syntactic context 1, but yields an ungrammaticalstring in context 2.

The first postulate would be that different types of ignitions should occur:

primed ignition in the case of the congruent sentence, but full unprimed

ignition after the item has been placed incorrectly. The activation of the

unprimed word web should give rise to a more substantial activity increasecompared to the primed ignition in the grammatical context. Therefore,there should be evidence of stronger activity increase in the cortex when a

word is placed in a context in which it is syntactically incorrect compared to

a situation in which it is positioned in a grammatically correct context. This

difference should be apparent early, shortly after the information about the

new word has reached the brain.

Page 284: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 284/331

13.2 Neuronal Grammar and Syntax-Related Brain Potentials 267

Neurophysiological studies using electroencephalography (EEG) show

that certain syntactically deviant word strings elicit an early negative event-

related potential (Neville, Nicol, Barss, Forster, & Garrett, 1991). This brainresponse has frequently been found to be maximal over the left hemispheredominant for language. It can be elicited by the incorrectly placed word of 

in sentence (1), but other syntactically incorrect strings may also produce it.

(1) The scientist criticized Max’s of proof the theorem.

The brain response is sometimes called the early left-anterior negativity. In

this framework, it would be interpreted as a physiological indicator of the fullignition of the word web of a word placed incorrectly. The early negativity is

less pronounced or absent if words are presented in a grammatical sentence.The effect could be reproduced repeatedly for different types of syntactic

violations, and with some variation of the cortical topography and latency of 

the brain response ( Friederici, 1997; Friederici & Mecklinger, 1996; Gunter,

Friederici, & Schriefers, 2000; Hahne & Friederici, 1999). An early left-

anterior negativity was also reported with incorrectly placed inflectional

morphemes as well (Weyerts et al., 1997).The second postulate differential brain responses elicited by grammatical

and ungrammatical strings is as follows: The threshold control mechanism

becomes active only after full ignition. Thus, activity decrease would be pre-

dicted after a strong initial activation process when a syntactically incorrectly

placed word or morpheme occurs. Such pronounced activity decrease should

be absent when a grammatical word string is being processed.There is a second neurophysiological brain response that distinguishes

syntactically well-formed strings from deviant strings. It is a positive compo-nent of the event-related potential and is maximal over the back of the head.

Its latency is in the range of 600 ms after onset of the critical word, and is

therefore called the P600 component or the syntactic positive shift (Hagoort,Brown, & Groothusen, 1993; Osterhout & Holcomb, 1992). When an incor-

rectly placed word is being processed, the late positive shift can appear after

an earlier negative shift (Friederici & Mecklinger, 1996; Hahne & Friederici,

1999; Osterhout, 1997).

The suggestion derived from the neuronal grammar model is that themore positive electrical brain response to syntactic anomalies reflects theaction of the mechanism that reduces the level of cortical activity after a

strong activation process (full ignition of a word) has taken place.

The third postulate neuronal grammar allows for concerns the activa-

tion of sequence detectors and the spreading wave of activity following a

word that matches its syntactic context. This should again lead to stronger

Page 285: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 285/331

268 Neurophysiology of Syntax

Figure 13.1. Gender violations elicited both an early left-anterior negativity and a late posi-tive shift of the brain potential compared to sentences in which no such violation was present.Example stimuli are the German sentences “Sie bereiste das Land” and “Sie bereiste denLand,” in which the determiner in italics is syntactically anomalous. (Example in English: “Shetravels the land,” in which the article is either in the correct neuter case or in an incorrectcase, masculine.) The dotted lines show responses to violations, the bold lines responsesto the grammatical sentences. Reprinted with permission from Gunter, T. C., Friederici,A. D., & Schriefers, H. (2000). Syntactic gender and semantic expectancy: ERPs reveal early

autonomy and late interaction. Journal of Cognitive Neuroscience, 12 (4), 556–68.

activation processes when a grammatical string is being processed in com-

parison to the processing of a string including a syntactic violation. At this

point, there does not appear to be evidence for this claim. It should, in princ-

iple, be possible to find evidence for this, because if the model is correct, thebackward spreading of activity would be a longer lasting process – in partic-

ular, if the words are placed in the context of an elaborate sentence.

Concern may arise regarding the latencies at which these syntacticallyrelated brain responses have been found. Responses that occur more than

one-half second after the information about a word is present in the input

may appear unlikely to reflect the earliest aspects of language processes.After 500 ms, it is, in many cases, already possible to respond to a stimulus

word, and a brain response occurring at such long latencies is not likely to

reflect the processes at the heart of the language machinery (cf. Lehmann

& Skrandies, 1984; Pulvermuller, 1999b; Skrandies, 1998).

There are, however, at least two responses to this concern. First, the max-imal response can sometimes be seen late, but this does not exclude thepossibility that earlier differences are also present. This becomes apparent

from Figure 13.1, in which the late positivity may appear to onset substan-

tially before its peak. Second, it is possible that the use of new recording

techniques makes it possible to record much earlier onsets of the relevant

differences. Many early brain responses are smaller than the late responses

Page 286: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 286/331

13.2 Neuronal Grammar and Syntax-Related Brain Potentials 269

appearing approximately one-half second after the stimulus. Because many

early responses are small, they are more likely to be masked by noise. For

recording earlier dynamics related to the processing of words and serial-order information, it may be crucial to introduce techniques that allow for

improving the signal-to-noise ratio. This may make earlier brain dynamicsof syntactic processing more visible.

This brief chapter is intended to show that what could appear as a purely

theoretical proposal of a neuronal grammar can actually be connected to

neurophysiological data. The early syntactic negativity may reflect the un-

primed full ignition of the neuronal set of a word placed in a syntactically

deviant environment. The late positive shift may reflect activation of thethreshold regulation device caused by the substantial activity increase pro-

duced by the unprimed ignition. Further important neurophysiological phe-

nomena realted to grammatical processing (e.g., Kluender & Kutas, 1993;

Muller, King, & Kutas, 1997) are still awaiting models at the neuronal level

that may provide tentative neuroscientific explanations for them. Clearly,

future research and physiological research must show which aspects of themodel presented in Chapters 10–12 can be maintained and which must be

replaced. Predictions appear obvious, and tests are available.

Page 287: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 287/331

CHAPTER FOURTEEN

Linguistics and the Brain

Linguistics is the study of language. Language is a system of brain circuits.

To qualify the latter statement, one may cite the father of modern linguistics,

Ferdinand de Saussure, who claimed that language (i.e., the language system,

or langue) is a “concrete natural object seated in the brain” (de Saussure,1916). If linguistics is the study of language and language is in one sense a

system of brain circuits, one would expect linguists to be open to the study of brain circuits. However, the dominating view in linguistics appears to be that

language theories must be formulated in an abstract manner, not in terms

of neuron circuits. Section 14.1 asks why linguists favor abstract rather thanneuron-based formulations of language mechanisms. Section 14.2 discusses

a few thoughts about how an abstract theory of language may profit from a

brain basis.

14.1 Why Are Linguistic Theories Abstract?

As mentioned, the dominating view in linguistics is that language theories

must be formulated in an abstract way. This could be a trivial claim becauseit is clear that every scientific theory must include abstract concepts. How-

ever, this is not the point. The point is that abstract in this context excludes

explicit reference to the organic basis of the processes described in an ab-

stract fashion. Linguistic theory is abstract in the sense that it does not refer

to neurons. Why is this so?For a scientist, this may be difficult to understand. An astronomer work-

ing on star formation would probably be open to discussing molecule clouds

that can be inferred from the recordings performed (e.g., with a radio tele-

scope). The “linguistic mentality,” so to speak, transformed to astronomy

would result in a scholar who studies stars but refuses to speak about their

component substances and driving forces. The scholar may claim that stars

270

Page 288: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 288/331

14.1 Why Are Linguistic Theories Abstract? 271

should be discussed only in terms of abstract concepts, not in terms of gases

and their temperature.

The position that linguistic theories must be abstract appears to be in needof justification, which could be either a principled statement or an excuse.

One may argue that it is unreasonable to specify neuron circuits of lan-guage – as there was no hope for a nineteenth-century astronomer to find out

about the physical processes occurring in the center of the sun that cause

emission of light. One may posit that there is simply not enough knowl-

edge about brain processes – or, 100 years ago, about the interior of the

sun. However, given the immense knowledge about how the brain processes

language accumulated in the last 20 years or so, it appears more likely thatrelevant knowledge is already available for clarifying mechanisms of lan-guage in brain terms. At least, the tools appear to be available for obtaining

the crucial knowledge about the underlying physiological processes, given

there are theoretically crucial predictions.

What is necessary, then, are ideas about how to connect the level of lan-

guage description to that of the description of neurons. Piling up more neuro-

physiological and imaging data may not help much in this enterprise. Empir-

ical facts do not by themselves form a theory about the generation of sunlightor language. Theoretical work is required in the first place. The theoretical

efforts can lead to the generation of predictions that can be addressed in

crucial experiments. Lack of empirical data is never a very good excuse for

postponing the necessary theoretical labor.A reasonable excuse for not addressing language in brain terms may take

the following possibilities into consideration. Linguists may have difficulty

understanding the language of neuroscientists, and, conversely, neurosci-

entists may have difficulty understanding linguistic terminology. After all,the distance between linguistics and neuroscience is not smaller than that

between physics and chemistry. Given that such mutual comprehension dif-ficulty is relevant, the important problem may be a problem of translation

(Schnelle, 1996a, 1996c). Therefore, it may appear relevant to provide trans-

lations between the language of linguistic algorithms and that of nerve cells,

their connections, and their activity states. Again, a good deal of theoretical

work is required.

In one publication, Chomsky (2000) offers another reason for not talkingabout neuronal mechanisms of language. “It may well be that the relevant

elements and principles of brain structure have yet to be discovered” (p. 25).

However, brain-theoretical concepts referred to by expressions such as cell 

assembly, synfire chain, and memory cell are available, and some of the po-

tentially relevant principles of brain structure and function are well under-

stood and have been available for theorizing about language for some time,

Page 289: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 289/331

272 Linguistics and the Brain

although language researchers have only very recently been taking advan-

tage of them. However, Chomsky is partly correct. Developing new concepts

may be necessary when theorizing about brain mechanisms of language. Themain effort undertaken in this book is to propose and sharpen concepts –

for example, those labeled neuronal set and multiple reverberation.Insummary,itseemsthereisnogoodreasonwhylinguistictheoriesshould

necessarily be abstract, and why they should not be formulated in concrete

terms – for example, in terms of neurons. Rather, this appears to be a crucial

linguistic task of the future.

14.2 How May Linguistic Theory Profit from a Brain Basis?

A question frequently asked by linguists is what the study of the brain couldbuy them. Clearly, if one is interested in language, one need not be interested

in the brain as well. There is absolutely no problem with this position. A

problem may arise, however, if one wishes to understand language as a

biological system, or as a “brain organ,” and still refuses to become concrete

about brain processes and representations (see Chomsky, 1980). In this case,

it might be advantageous to attempt to connect one’s terminology to theputative mechanisms.

It is possible that translation issues can be solved and a language can be

developed that refers to linguistic structures and brain structures, linguistic

processes and brain processes, and to underlying linguistic principles, as well

as to neuroscientific principles of brain structure and function. Given sucha language is available, it would be possible to explore the space of possi-

bilities that is restricted on the one side by current neuroscientific knowl-

edge and on the other side by linguistic phenomena. Using neuroscientificknowledge and data for guiding linguistic theorizing appears to be fruitful.

Testing syntax theory in neurophysiological experiments may be fruitful as

well. Thus, neuroscientific data could then constrain linguistic theory. Thereverse may also be true. The study of language may, given such a language

connecting noun phrases to neurons (Marshall, 1980) is available, allow for

making predictions on brain circuits that have not been detected by other

means. Availability of a brain–language interface of this type, a neuronal

language theory, may be a necessary condition for deciding between alter-native approaches to grammar as it could be a tool for exploring neuron

circuits specific to the human brain. A language theory at the neuronal level

is required in cognitive neuroscience.

How would the situation, when performing brain imaging studies of lan-

guage, improve if a brain–language interface were available? It may appear

that abstract linguistic theories have some implications for neuroscientific

Page 290: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 290/331

14.2 How May Linguistic Theory Profit from a Brain Basis? 273

structures and processes. One may therefore claim that interesting questions

about brain–language relationships can already be investigated in neurosci-

entific experiments without additional theoretical work in the first place. Theargument here is that this is not true. Linguistic theories do not in fact have

strong implications regarding neuroscientific facts, and the experimenternecessarily ends up in difficulties if he or she wishes to entertain them as

research guidelines.

A rather fresh view on the relation of neuroscience and linguistics has

been put forward by Chomsky (1980) in his book, Rules and Representa-

tions. Chomsky considered how the presence of an invisible representation

of a certain kind during the processing of a sentence could be tested in anelectrophysiological experiment, taking as an example the “wh- sentence

structure.” Linguists assign such structures to one type of sentence, type A.

Now there is a second category of sentences, type B, for which such assign-

ment is implied by one of Chomsky’s theories, but not by more traditional

approaches to syntax. His proposal about the role of a neurophysiological

experiment in this theoretical issue is as follows: If “a certain pattern of electrical activity” is highly correlated with the clear cases in which a wh-

sentence is processed (type A sentences), and if this same neuronal patternalso occurs during processing of a sentence for which such a representation

has been postulated based on theoretical considerations (type B sentence),

then one would have evidence that this latter representation of the sentence

in question is psychologically and biologically real.Therefore, in principle, there appears to be no problem with the neurosci-

entific test of linguistic theories. However, Chomsky’s view is not realistic.

A closer look at the actual empirical data obtained so far, indicates that

clear correlation between language phenomena and patterns of electricalactivity are not easy to find. Recent studies of syntactic phenomena have

great difficulty in proving that the physiological phenomena that are re-ported to co-occur with linguistic properties of sentences are strictly related

to these linguistic properties per se. As an alternative, they may be related

to psychological phenomena in which a linguist is probably less interested,

such as working memory load and retrieval (see Kluender & Kutas, 1993).

Thus, there would not be a high correlation between a bold brain response

and a well-defined linguistic process, but there would be a brain responsecooccurring with a variety of linguistic and psychological phenomena.

But this is only the first reason why Chomsky’s view is not appropriate.

A further important theoretical problem that a physiological test of lin-

guistic ideas must face – ignored by Chomsky – is as follows: The instru-

ments for monitoring brain activity do not by themselves tell the researcherwhat to look for when investigating linguistic representations and processes.

Page 291: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 291/331

274 Linguistics and the Brain

There are infinite possibilities for describing and analyzing a short time series

obtained, for instance, using a multichannel electro- or magnetoencephalo-

graph. What should the language and brain scientist begin with when search-ing for the pattern that, in the clear case, correlates with the occurrence of 

a wh- sentence? Answers to this question can be provided only by a theoryabout the brain mechanisms of language.

A good deal of progress being made in the language and brain sciences is

exactly of this nature: finding out what the relevant patterns of brain activity

might be. Once again, it must be stressed that brain theory reflections were

the seed of such progress.

To mention but one example, it has been speculated by a scholar whofurther developed the Hebbian brain theory (Milner, 1974) that fast oscil-

lations in the brain may be related to the activity of neuronal ensembles

(see also Freeman, 1975; von der Malsburg, 1985). Fast rhythmic brain ac-

tivity may result from reverberation of neuronal activity in loops of cortical

connections that conduct activity in the millisecond range (Chapter 8), or

they may emerge from interaction between cell assembly neurons and theirinhibitory neighbors (cf. Chapter 2). The idea that high-frequency brain ac-

tivity is related to the activation of neuronal ensembles and that it may evenbe a correlate of perceptual and higher cognitive processes inspired nu-

merous experimental investigations in the neurosciences. The results were

largely supportive (Bressler & Freeman, 1980; Singer & Gray, 1995). It was

therefore proposed that the cell assemblies, or functional webs, representingwords produce high-frequency responses when activated. On the basis of this

theoretically motivated speculation, several experiments have been con-

ducted, all of which converge on the conclusion that dynamics in high-

frequency cortical responses distinguish word forms from meaningless pseu-dowords (see Chapter 4 for details). More important, the high-frequency

responses related to word processes frequently exhibited a specific topo-graphy not revealed in other studies of high-frequency responses of cognitive

processes (Pulvermuller et al., 1997; Tallon-Baudry & Bertrand, 1999). Fur-

thermore, high-frequency brain responses recorded from motor and

visual areas distinguished between words from different categories, and may

thus reveal elementary information about more specific properties of word

processing in the brain (Pulvermuller, Keil, & Elbert, 1999; Pulvermuller,Lutzenberger et al., 1999).

In the present context, the important conclusion may be that patterns

of brain activity possibly indicating aspects of linguistic processing were

discovered only because educated guesses about the brain basis of word

processing were possible. Without the theorizing these guesses are built on,the probability of finding the patterns would have approximated zero. This

Page 292: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 292/331

14.2 How May Linguistic Theory Profit from a Brain Basis? 275

is the second reason why Chomsky’s idea about a neurobiological test of 

linguistic constructs is unrealistic: He ignores the biotheoretical prerequisites

for such testing.

It can be concluded that although leading linguists may see the possible

fruitfulness of neuroscientific research into language, the nature of the major

tasks to be addressed in the theoretical neuroscience of language is largely

neglected. For serious empirical investigation of the brain mechanisms of 

language, it is not enough to provide abstract descriptions of language phe-

nomena; it is also necessary to spell out possible language mechanisms in

terms of neuronal circuitry.

Building a theory of brain mechanism of language is certainly not an easytask. It is clear that such theorizing will not immediately result in the ulti-

mate answer to the relevant language-related questions. Most likely, early

proposals will be falsified by experimental brain research – as, for example,

Helmholtz’s idea that “Gravitationsenergie” causes the sun to shine had to

be replaced. However, the important point at this stage may be to make

theoretically relevant brain research on language possible. Scientific inves-

tigation of the interesting questions in linguistics requires a brain model of 

the relevant linguistic processes. The purpose of this book is to give it a try.

Page 293: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 293/331

Page 294: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 294/331

References

Abeles, M. (1991). Corticonics – Neural circuits of the cerebral cortex. Cambridge:Cambridge University Press.

Abeles, M., Bergman, H., Gat, I., Meilijson, I., Seidemann, E., Tishby, N., & Vaadia, E.(1995). Cortical activity flips among quasi-stationary states. Proceedings of the

National Academy of Sciences, USA, 92, 8616–20.Abeles, M., Bergman, H., Margalit, E., & Vaadia, E. (1993). Spatiotemporal firing

patterns in the frontal cortex of behaving monkeys. Journal of Neurophysiology,70, 1629–38.

Aboitiz, F., Scheibel, A. B., Fisher, R. S., & Zaidel, E. (1992). Fiber composition of the human corpus callosum. Brain Research, 598, 143–53.

Ahissar, E., Vaadia, E., Ahissar, M., Bergman, H., Arieli, A., & Abeles, M. (1992).Dependence of cortical plasticity on correlated activity of single neurons and onbehavior context. Science, 257 , 1412–15.

Ajukiewicz, K. (1936). Die syntaktische Konnexitat. Studia Philosophica, 1, 1–27.Angrilli, A., Dobel, C., Rockstroh, B., Stegagno, L., & Elbert, T. (2000). EEG brain

mapping of phonological and semantic tasks in Italian and German languages.Clinical Neurophysiology, 111, 706–16.

Assadollahi, R., & Pulvermuller, F. (2001). Neuromagnetic evidence for early accessto cognitive representations. Neuroreport, 12, 207–13.

Bach, E., Brown, C., & Marslen-Wilson, W. (1986). Crossed and nested dependenciesin German and Dutch: a psycholinguistic study. Language and Cognitive Processes,

1, 249–62.Bak, T. H., O’Donovan, D. G., Xuereb, J. H., Boniface, S., & Hodges, J. R. (2001).

Selective impairment of verb processing associated with pathological changes in

Brodmann areas 44 and 45 in the motor neuron disease–dementia–aphasia syn-drome. Brain, 124, 103–20.

Baker, G. P., & Hacker, P. M. S. (1984). Language, sense and nonsense. Oxford: BasilBlackwell.

Bar-Hillel, Y., Perles, M., & Shamir, E. (1961). On formal properties of simple phrasestructure grammars. Zeitschrift f   ur Phonetik, Sprachwissenschaft und Kommunika-

tionsforschung, 14, 143–72.

277

Page 295: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 295/331

Page 296: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 296/331

References 279

Bressler, S. L., & Freeman, W. J. (1980). Frequency analysis of olfactory system EEGin cat, rabbit and rat. Electroencephalography and Clinical Neurophysiology, 50,19–24.

Broca, P. (1861). Remarques sur la siege de la facult´ e de la parole articul´ ee, suiviesd’une observation d’aph´ emie (perte de parole). Bulletin de la Soci´ et   e d’Anatomie,

 36, 330–57.Brodmann, K. (1909). Vergleichende Lokalisationslehre der Großhirnrinde. Leipzig:

Barth.Brown, C. M., Hagoort, P., & ter Keurs, M. (1999). Electrophysiological signatures

of visual lexical processing: open- and closed-class words. Journal of Cognitive

Neuroscience, 11, 261–81.Brown, T. H., Kairiss, E. W., & Keeman, C. L. (1996). Hebbian synapses: biophysical

mechanisms and algorithms. Review in Neuroscience, 13, 475–511.Brown, W. S., & Lehmann, D. (1979). Verb and noun meaning of homophone words

activate different cortical generators: a topographic study of evoked potentialfields. Experimental Brain Research, 2, S159–68.

Bryden, M. P., Hecaen, H., & DeAgostini, M. (1983). Patterns of cerebral organiza-tion. Brain and Language, 20(2), 249–62.

Buonomano, D. V. (2000). Decoding temporal information: A model based on short-term synaptic plasticity. J Neuroscience, 20, 1129–41.

Buonomano, D. V., & Merzenich, M. M. (1998). Cortical plasticity: from synapses to

maps. Annual Review of Neuroscience, 21, 149–86.Caplan, D. (1987). Neurolinguistics and linguistic aphasiology. An introduction.

Cambridge, MA: Cambridge University Press.Caplan, D. (1996). Language: structure, processing, and disorders. Cambridge, MA:

MIT Press.Cappa,S.F.,Binetti,G.,Pezzini,A.,Padovani,A.,Rozzini,L.,&Trabucchi,M.(1998).

Object and action naming in Alzheimer’s disease and frontotemporal dementia.Neurology, 50(2), 351– 5.

Caramazza, A., & Zurif, E. B. (1976). Dissociation of algorithmic and heuristic pro-

cesses in sentence comprehension: evidence from aphasia.Brain and Language, 3,572–82.

Charniak, E. (1993). Statistical language learning. Cambridge, MA: MIT Press.Cheour, M., Ceponiene, R., Lehtokoski, A., Luuk, A., Allik, J., Alho, K., & Naatanen,

R. (1998). Development of language-specific phoneme representations in the in-fant brain. Nature Neuroscience, 1, 351–3.

Chomsky, N. (1957). Syntactic structures. The Hague: Mouton.Chomsky, N. (1963). Formal properties of grammars. In R. D. Luce & R. R. Bush &

E. Galanter (Eds.), Handbook of mathematical psychology, Vol. 2 (pp. 323–418).

New York, London: Wiley.Chomsky, N. (1965). Aspects of the theory of syntax. Cambridge, MA: MIT Press.Chomsky, N. (1980). Rules and representations. New York: Columbia University

Press.Chomsky, N. (2000). New horizons in the study of language and mind. Cambridge,

MA: Cambridge University Press.

Page 297: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 297/331

280 References

Creutzfeldt, O., Ojemann, G., & Lettich, E. (1989). Neuronal activity in the humanlateral temporal lobe. I. Responses to speech. Experimental Brain Research, 77 ,451–75.

Cruse, H., Bartling, C., Cymbalyuk, G., Dean, J., & Dreifert, M. (1995). A modularartificial neural net for controlling a six-legged walking system. Biological Cyber-

netics, 72, 421–30.Cruse, H., & Bruwer, M. (1987). The human arm as a redundant manipulator: the

control of path and joint angles. Biological Cybernetics, 57 , 137–44.Dale, A. M., Liu, A. K., Fischl, B. R., Buckner, R. L., Belliveau, J. W., Lewine, J. D.,

& Halgren, E. (2000). Dynamic statistical parametric mapping: combining fMRIand MEG for high-resolution imaging of cortical activity. Neuron, 26(1), 55–67.

Damasio, A. R. (1989). The brain binds entities andeventsby multiregional activation

from convergence zones. Neural Computation, 1, 123–32.Damasio, H., Grabowski, T. J., Tranel, D., Hichwa, R. D., & Damasio, A. R. (1996).

A neural basis for lexical retrieval. Nature, 380, 499–505.Damasio, A. R., & Tranel, D. (1993). Nouns and verbs are retrieved with differently

distributed neural systems. Proceedings of the National Academy of Sciences, USA,

90, 4957–60.Daniele, A., Giustolisi, L., Silveri, M. C., Colosimo, C., & Gainotti, G. (1994). Ev-

idence for a possible neuroanatomical basis for lexical processing of nouns andverbs. Neuropsychologia, 32, 1325–41.

Deacon, T. W. (1992). Cortical connections of the inferior arcuate sulcus cortex inthe macaque brain. Brain Research, 573(1), 8–26.

DeFelipe, J., & Farinas, I. (1992). The pyramidal neuron of the cerebral cortex:morphological and chemical characteristics of the synaptic inputs. Progress in

Neurobiology, 39(6), 563–607.Dehaene, S., Changeux, J. P., & Nadal, J. P. (1987). Neural networks that learn tem-

poral sequences by selection. Proceedings of the National Academy of Sciences,

USA, 84, 2727–31.Dehaene-Lambertz, G., & Dehaene, S. (1994). Speed and cerebral correlates of 

syllable discrimination in infants. Nature, 370, 292–5.Dell, G. S. (1986). A spreading-activation theory of retreival in sentence production.

Psychological Review, 93, 283–321.Dell, G. S., Schwartz, M. F., Martin, N., Saffran, E. M., & Gagnon, D. A. (1997).

Lexical access in aphasic and nonaphasic speakers. Psychological Review, 104(4),801–38.

De Renzi, E., & Vignolo, L. (1962). The Token Test: a sensitive test to detect receptivedisturbances in aphasics. Brain, 85, 665–78.

de Saussure, F. (1916). Cours de linguistique generale. Paris: Payot.

Devlin, J. T., Russell, R. P., Davis, M. H., Price, C. J., Moss, H. E., Fadili, M. J., &Tyler, L. K. (2002). Is there an anatomical basis for category-specificity? Semanticmemory studies in PET and fMRI. Neuropsychologia, 40, 54–75.

Diesch, E., Biermann, S., & Luce, T. (1998). The magnetic mismatch field elicited bywords and phonological non-words. NeuroReport, 9, 455–60.

Diesmann, M., Gewaltig, M. O., & Aertsen, A. (1999). Stable propagation of syn-chronous spiking in cortical neural networks. Nature, 402(6761), 529–33.

Page 298: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 298/331

References 281

Dilthey, W. (1989). Einleitung in die Geisteswissenschaften [Introduction to the human

 sciences], Vol. 1. Princeton, NJ: Princeton University Press.Dobel, C., Pulvermuller, F., Harle, M., Cohen, R., Kobbel, P., Schonle, P. W., &

Rockstroh, B. (2001). Syntactic and semantic processing in the healthy and aphasichuman brain. Experimental Brain Research, 140, 77–85.

Egelhaaf, M., Borst, A., & Reichardt, W. (1989). Computational structure of a bio-logical motion-detection system as revealed by local detector analysis in the fly’snervous system. Journal of the Optical Society of America (A), 6, 1070–87.

Eisenberg, P. (1999). Grundriss der deutschen Grammatik: Der Satz, Vol. 2. Stuttgart:Verlag J.B. Metzler.

Elbert, T., Pantev, C., Wienbruch, C., Rockstroh, B., & Taub, E. (1995). Increasedcortical representation of the fingers of the left hand in string players. Science,

 270(5234), 305–7.Ellis, A. W., & Young, A. W. (1988). Human cognitive neuropsychology. Hove, UK:

Lawrence Erlbaum Associates Ltd.Elman, J. L. (1990). Finding structure in time. Cognitive Science, 14, 179–211.Elman, J. L., Bates, L., Johnson, M., Karmiloff-Smith, A., Parisi, D., & Plunkett, K.

(1996). Rethinking innateness: a connectionist perspective on development .Cambridge, MA: MIT Press.

Engert, F., & Bonhoeffer, T. (1999). Dendritic spine changes associated with hip-pocampal long-term synaptic plasticity. Nature, 399(6731), 66–70.

Epstein, H. T. (1999). Other brain effects of words. Behavioral and Brain Sciences, 22, 287–8.

Eulitz, C., Eulitz, H., Maess, B., Cohen, R., Pantev, C., & Elbert, T. (2000). Magneticbrain activity evoked and induced by visually presented words and nonverbalstimuli. Psychophysiology, 37 (4), 447–55.

Farah, M. J., & McClelland, J. L. (1991). A computational model of semantic mem-ory impairment: modality specificity and emergent category specificity. Journal of 

Experimental Psychology: General, 120, 339–57.Federmeier, K. D., Segal, J. B., Lombrozo, T., & Kutas, M. (2000). Brain responses

to nouns, verbs and class-ambiguous words in context. Brain, 123, 2552–66.Fiez, J. A., & Petersen, S. E. (1998). Neuroimaging studies of word reading. Proceed-

ings of the National Academy of Science, USA, 95(3), 914–21.Fiez, J. A., Raichle, M. E., Balota, D. A., Tallal, P., & Petersen, S. E. (1996). PET

activation of posterior temporal regions during auditory word presentation andverb generation. Cerebral Cortex, 6, 1–10.

Freeman, W. J. (1975). Mass action in the nervous system. New York: Academic Press.Freud, S. (1891). Zur Auffassung der Aphasien. Leipzig, Wien: Franz Deuticke.Friederici, A., & Mecklinger, A. (1996). Syntactic parsing as revealed by brain re-

sponses: first pass and second pass parsing processes. Journal of PsycholinguisticResearch, 25, 157–76.

Friederici, A. D. (1997). Neurophysiological aspects of language processing.Clinical 

Neuroscience, 4, 64–72.Friederici, A. D., Pfeifer, E., & Hahne, A. (1993). Event-related brain potentials

during natural speech processing: effects of semantic, morphological and syntacticviolations. Cognitive Brain Research, 1, 183–92.

Page 299: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 299/331

282 References

Fry, D. B. (1966). The development of the phonological system in the normal anddeaf child. In F. Smith & G. A. Miller (Eds.), The genesis of language (pp. 187–206).Cambridge, MA: MIT Press.

Fuster, J. M. (1995). Memory in the cerebral cortex. An empirical approach to neural networks in the human and nonhuman primate. Cambridge, MA: MIT Press.

Fuster, J. M. (1997). Network memory. Trends in Neurosciences, 20, 451–9.Fuster, J. M. (1998a). Distributed memory for both short and long term.Neurobiology

in Learning and Memory, 70(1–2), 268–74.Fuster, J. M. (1998b). Linkage at the top. Neuron, 21(6), 1223–4.Fuster, J. M. (1999). Hebb’s other postulate at work on words. Behavioral and Brain

Sciences, 22, 288–9.Fuster, J. M. (2000). Cortical dynamics of memory. International Journal of 

Psychophysiology, 35(2–3), 155–64.Fuster, J. M., & Alexander, G. E. (1971). Neuron activity related to short-term mem-

ory. Science, 173(997), 652–4.Fuster, J. M., Bodner, M., & Kroger, J. K. (2000). Cross-modal and cross-temporal

association in neurons of frontal cortex. Nature, 405(6784), 347–51.Fuster, J. M., & Jervey, J. P. (1982). Neuronal firing in the inferiotemporal cortex of 

the monkey in a visual memory task. Journal of Neuroscience, 2, 361–75.Gaifman, C. (1965). Dependency systems and phrase structure systems.Information

and Control, 8, 304–37.

Galuske, R. A., Schlote, W., Bratzke, H., & Singer, W. (2000). Interhemispheric asym-metries of the modular structure in human temporal cortex. Science, 289(5486),1946–9.

Garrett, M. (1980). Levels of processing in sentence production. In B. Butterworth(Ed.), Language Production I (pp. 177–220). London: Academic Press.

Gazdar, G., Klein, E., Pullum, G., & Sag, I. (1985). Generalized phrase structure

 grammar . Cambridge, MA: Harvard University Press.Geschwind, N. (1965a). Disconnection syndromes in animals and man (1).Brain, 88,

237–94.

Geschwind, N. (1965b). Disconnection syndromes in animals and man (2).Brain, 88,585–644.

Geschwind,N. (1974).Selectedpapersonlanguageandthebrain. Dordrecht: Kluewer.Geschwind, N., & Levitsky, W. (1968). Human brain: left–right asymmetries in tem-

poral speech region. Science, 161, 186–7.Glaser, W. R., & Dungelhoff, F. J. (1984). The time course of picture–word interfer-

ence. Journal of Experimental Psychology: Human Perception and Performance,

10, 640–54.Goodglass, H., & Quadfasel, F. A. (1954). Language laterality in left-handed aphasics.

Brain, 77 , 521–48.Grabowski, T. J., Damasio, H., & Damasio, A. R. (1998). Premotor and pre-

frontal correlates of category-related lexical retrieval. Neuroimage, 7 , 232–43.

Gunter, T. C., Friederici, A. D., & Schriefers, H. (2000). Syntactic genderand semanticexpectancy: ERPs reveal early autonomy and late interaction. Journal of Cognitive

Neuroscience, 12(4), 556–68.

Page 300: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 300/331

References 283

Haegeman, L. (1991). Introduction to government and binding theory. Cambridge,MA: Basil Blackwell.

Hagoort, P., Brown, C., & Groothusen, J. (1993). The syntactic positive shift (SPS)

as an ERP-measure of syntactic processing. Language and Cognitive Processes,8, 439–483.

Hagoort, P., Indefrey, P., Brown, C., Herzog, H., Steinmetz, H., & Seitz, R. J. (1999).The neural circuitry involved in the reading of German words and pseudowords:a PET study. Journal of Cognitive Neuroscience, 11(4), 383–98.

Hahne, A., & Friederici, A. D. (1999). Electrophysiological evidence for two stepsin syntactic analysis. Early automatic and late controlled processes. Journal of 

Cognitive Neuroscience, 11(2), 194–205.Hare, M., Elman, J. L., & Daugherty, K. G. (1995). Default generalisation in connec-

tionist networks. Language and Cognitive Processes, 10, 601–30.Harris, Z. S. (1945). Discontinuous morphemes. Language, 21, 121–7.Harris, Z. S. (1951). Structural linguistics. Chicago: Chicago University Press.Harris, Z. S. (1952). Discourse analysis. Language, 28, 1–30.Harris, Z. S. (1955). From phonemes to morphemes. Language, 31, 190–222.Harris, Z. S. (1957). Co-occurrence and transformation in linguistic structure. Lan-

 guage, 33, 283–340.Harris, Z. S. (1965). Transformational theory. Language, 41, 363–401.Hasbrooke, R. E., & Chiarello, C. (1998). Bihemispheric processing of redundant

bilateral lexical information. Neuropsychology, 12, 78–94.Hauk, O., & Pulvermuller, F. (2002). Neurophysiological distinction of action words

in the frontal lobe: an ERP study using minimum current estimates.Hayes, T. L., & Lewis, D. (1993). Hemispheric differences in layer III pyramidal

neurons of the anterior language area. Archives of Neurology, 50, 501–5.Hays, D. G. (1964). Dependency theory: a formalism and some observations.

Language, 40, 511–25.Hebb, D. O. (1949). The organization of behavior. A neuropsychological theory. New

York: John Wiley.

Hecaen, H., De Agostini, M., & Monzon-Montes, A. (1981). Cerebral organizationin left-handers. Brain and Language, 12(2), 261–84.

Hellwig, B. (2000). A quantitative analysis of the local connectivity between pyra-midal neurons in layers 2/3 of the rat visual cortex. Biological Cybernetics, 82(2),111–21.

Heringer, H. J. (1996). Deutsche Syntax dependentiell . Tubingen: StauffenburgVerlag.

Hetherington, P. A., & Shapiro, M. L. (1993). Simulating Hebb cell assemblies: thenecessity for partitioned dendritic trees and a post-not-pre LTD rule. Network, 4,

135–53.Hinton, G. E., & Shallice, T. (1991). Lesioning an attractor network: investigation of 

acquired dyslexia. Psychological Review, 98, 74–95.Hubel, D. (1995). Eye, brain, and vision (2 ed.). New York: Scientific American

Library.Humphreys, G. W., Evett, L. J., & Taylor, D. E. (1982). Automatic phonological

priming in visual word recognition. Memory and Cognition, 10, 576–90.

Page 301: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 301/331

Page 302: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 302/331

References 285

Kreiter, A. K. (2002). Functional implications of temporal structure in primate cor-tical information processing. Zoology: Analysis of Complex Systems, 104, 241–55.

Kujala, T., Alho, K., & Naatanen, R. (2000). Cross-modal reorganization of human

cortical functions. Trends in Neuroscience, 23(3), 115–20.Lambek, J. (1958). The mathematics of sentence structure. Americal Mathematical 

Monthly, 65, 154–70.Lambek, J. (1959). Contributions to a mathematical analysis of the English verb

phrase. Journal of the Canadian Linguistic Association, 5, 83–9.Landauer, T. K., & Dumais, S. T. (1997). A solution to Plato’s problem: the Latent

Semantic Analysis theory of acquisition, induction, and representation of knowledge. Psychological Review, 104, 211–40.

Lashley, K. S. (1950). In search of the engram. Symposium of the Society for Experi-

mental Biology, 4, 454–82.Lashley, K. S. (1951). The problem of serial order in behavior. In L. A. Jeffress (Ed.),

Cerebral mechanisms in behavior. The Hixxon symposium, pp. 112–36. New York:John Wiley.

Le Clec’H, G., Dehaene, S., Cohen, L., Mehler, J., Dupoux, E., Poline, J. B., Lehericy,S., van de Moortele, P. F., & Le Bihan, D. (2000). Distinct cortical areas for names of numbers and body parts independent of language and input modality.Neuroimage,

12, 381–91.Lee, K. H., Chung, K., Chung, J. M., & Coggeshall, R. E. (1986). Correlation of cell

body size, axon size, and signal conduction velocity for individually labelled dorsalroot ganglion cells in the cat. Journal of Comparative Neurology, 243(3), 335–46.

Lehmann, D., & Skrandies, W. (1984). Spatial analysis of evoked potentials in man:a review. Progress in Neurobiology, 23, 227–50.

Levelt, W. J. M. (1974). Formal grammars in linguistics and psycholinguistics. Vol. 1.

 An introduction to the theory of formal languages and automata. The Hague, Paris:Mouton.

Levelt, W. J. M., Roelofs, A., & Meyer, A. S. (1999). A theory of lexical access inspeech production. Behavioral and Brain Sciences, 22, 1–75.

Levelt, W. J., Schriefers, H., Vorberg, D., Meyer, A. S., Pechmann, T., & Havinga, J.(1991). The time course of lexical access in speech production: A study of picturenaming. Psychological Review, 98, 122–42.

Lichtheim, L. (1885). On aphasia. Brain, 7 , 433–84.Locke, J. L. (1989). Babbling and early speech: continuity and individual differences.

First Language, 9, 191–206.Locke, J. L. (1993). The child’s path to spoken language. Cambridge, MA: Harvard

University Press.Lutzenberger, W., Pulvermuller, F., & Birbaumer, N. (1994). Words and pseudowords

elicit distinct patterns of 30-Hz activity in humans. Neuroscience Letters, 176,115–18.

Marcel, A. J., & Patterson, K. (1978). Word recognition and production: reciprocityin clinical and normal studies. In J. Requin (Ed.), Attention and performance

(Vol. VII, pp. 209–26). New Jersey: Lawrence Erlbaum Associates.Marcus, S. (1965). Sur la Notion de la Projectivite.Zeitschrift f   ur mathematische Logik

und Grundlagen der Mathematik, 11, 181–92.

Page 303: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 303/331

286 References

Marie, P. (1906). Revision de la question de l’aphasie: la triosieme circonvolutionfrontale gauche ne joue uncun role special dans la fonction du langage. Semaine

Medicale (Paris), 26, 241–7.

Markov, A. A. (1913). Essai d’une recherche statistique sur le texte du roman“Eugene Onegin.” Bulletin de l’Academie Imperiale des Sciences, St. Petersburg, 7 .

Marshall, J. C. (1980). On the biology of language acquisition. In D. Caplan (Ed.),Biological studies of mental processes. Cambridge, MA: MIT Press.

Marslen-Wilson, W., & Tyler, L. (1997). Dissociating types of mental computation.Nature, 387 , 592–4.

Marslen-Wilson, W., & Tyler, L. K. (1975). Processing structure of sentence percep-tion. Nature, 257 (5529), 784–6.

Marslen-Wilson, W., Tyler, L. K., Waksler, R., & Older, L. (1994). Morphology and

meaning in the English mental lexicon. Psychological Review, 101, 3–33.Marslen-Wilson, W., & Warren, P. (1994). Levels of perceptual representation and

process in lexical access: words, phonemes, and features. Psychological Review,

101, 633–75.Marslen-Wilson, W. D., & Tyler, L. K. (1980). The temporal structure of spoken

language understanding. Cognition, 8, 1–71.Martin, A., & Chao, L. L. (2001). Semantic memory and the brain: structure and

processes. Current Opinion in Neurobiology, 11, 194–201.Martin, A., Wiggs, C. L., Ungerleider, L. G., & Haxby, J. V. (1996). Neural correlates

of category-specific knowledge. Nature, 379, 649–52.Mason, A., Nicoll, A., & Stratford, K. (1991). Synaptic transmission between indi-

vidual pyramidal neurons of the rat visual cortex in vitro. Journal of Neuroscience,

11(1), 72–84.McClelland, J. L., & Rumelhart, D. E. (1981). An interactive activation model of 

context effects in letter perception: I. An account of basic findings. Psychological 

Review, 88, 375–407.McCulloch, W. S., & Pitts, W. H. (1943). A logical calculus of ideas immanent in

nervous activity. Bulletin of Mathematical Biophysics, 5, 115–33.

Merzenich, M. M., Kaas, J. H., Wall, J., Nelson, R. J., Sur, M., & Felleman, D. (1983).Topographic reorganization of somatosensory cortical areas 3b and 1 in adultmonkeys following restricted deafferentation. Neuroscience, 8(1), 33–55.

Merzenich, M. M., Kaas, J. H., Wall, J. T., Sur, M., Nelson, R. J., & Felleman,D. J. (1983). Progression of change following median nerve section in the corticalrepresentation of the hand in areas 3b and 1 in adult owl and squirrel monkeys.Neuroscience, 10(3), 639–65.

Mesulam, M. M. (1990). Large-scale neurocognitive networks and distributedprocessing for attention, language, and memory. Annals of Neurology, 28,

597–613.Meyer, D. E., & Schvaneveldt, R. W. (1971). Facilitation in recognizing pairs of 

words: eviedence of a dependence of retrieval operations. Journal of Experimental 

Psychology, 90, 227–35.Miceli, G., Silveri, C., Nocentini, U., & Caramazza, A. (1988). Patterns of disso-

ciation in comprehension and production of nouns and verbs. Aphasiology, 2,351–8.

Page 304: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 304/331

References 287

Miceli, G., Silveri, M., Villa, G., & Caramazza, A. (1984). On the basis of agrammatics’difficulty in producing main verbs. Cortex, 20, 207–20.

Miller, G. A., & Chomsky, N. (1963). Finite models of language users. In R. D.

Luce & R. R. Bush & E. Galanter (Eds.), Handbook of mathematical psychology,Vol. 2 (pp. 419–91). New York, London: Wiley.

Miller, R. (1991). Cortico-hippocampal interplay and the representation of contexts

in the brain. Berlin: Springer.Miller, R. (1996). Axonal conduction times and human cerebral laterality. A psychobi-

ological theory. Amsterdam: Harwood Academic Publishers.Miller, R., & Wickens, J. R. (1991). Corticostriatal cell assemblies in selective at-

tention and in representation of predictable and controllable events: a generalstatement of corticostriatal interplay and the role of striatal dopamine. Concepts

in Neuroscience, 2, 65–95.Milner, P. M. (1957). The cell assembly: Mk. II. Psychological Review, 64, 242–52.Milner, P. M. (1974). A model for visual shape recognition. Psychological Review,

81, 521–35.Milner, P. M. (1996). Neural representation: some old problems revisited.Journal of 

Cognitive Neuroscience, 8, 69–77.Minsky, M. (1972). Computation: finite and infinite machines. London: Prentice-Hall.Minsky, M., & Papert, S. (1969). Perceptrons. Cambridge, MA: MIT Press.Mohr, B., & Pulvermuller, E. (2002). Redundancy gains and costs in word processing:

the effect of short stimulus onset asynchronies (SOAs). Journal of Experimental Psychology: Learning, Memory and Cognition.

Mohr, B., Pulvermuller, F., Mittelstadt, K., & Rayman, J. (1996). Multiple stimuluspresentation facilitates lexical processing. Neuropsychologia, 34, 1003–13.

Mohr, B., Pulvermuller, F., Rayman, J., & Zaidel, E. (1994). Interhemispheric co-operation during lexical processing is mediated by the corpus callosum: evidencefrom the split-brain. Neuroscience Letters, 181, 17–21.

Mohr, B., Pulvermuller, F., & Zaidel, E. (1994). Lexical decision after left, right andbilateral presentation of content words, function words and non-words: evidence

for interhemispheric interaction. Neuropsychologia, 32, 105–24.Molfese, D. L. (1972). Cerebral asymmetry in infants, children and adults: auditory

evoked responses to speech and noise stimuli. Unpublished doctoral dissertation:Pennsylvania State University.

Molfese, D. L., Burger-Judisch, L. M., Gill, L. A., Golinkoff, R. M., & Hirsch-Pasek,K. A. (1996). Electrophysiological correlates of noun-verb processing in adults.Brain and Language, 54, 388–413.

Montoya, P., Larbig, W., Pulvermuller, F., Flor, H., & Birbaumer, N. (1996). Corticalcorrelates of semantic classical conditioning. Psychophysiology, 33, 644–49.

Moore, C. J., & Price, C. J. (1999). A functional neuroimaging study of the variablesthat generate category-specific object processing differences. Brain, 122, 943–62.

Moro, A., Tettamanti, M., Perani, D., Donati, C., Cappa, S. F., & Fazio, F. (2001).Syntax and the brain: disentangling grammar by selective anomalies.Neuroimage,

13, 110–18.Morton, J. (1969). The interaction of information in word recognition. Psychological 

Review, 76, 165–78.

Page 305: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 305/331

288 References

Muller, H. M., King, J. W., & Kutas, M. (1997). Event-related potentials elicited byspoken relative clauses. Cognitive Brain Research, 5, 193–203.

Mummery, C. J., Patterson, K., Hodges, J. R., & Price, C. J. (1998). Functional neu-

roanatomy of the semantic system: divisible by what? Journal of Cognitive Neuro- science, 10, 766–77.

Naatanen, R. (2001). The perception of speech sounds by the human brain as re-flected by the mismatch negativity (MMN) and its magnetic equivalent (MMNm).Psychophysiology, 38, 1–21.

Naatanen, R., Lehtokoski, A., Lennes, M., Cheour, M., Huotilainen, M., Iivonen, A.,Valnio, A., Alku, P., Ilmoniemi, R. J., Luuk, A., Allik, J., Sinkkonen, J., & Alho, K.(1997). Language-specific phoneme representations revealed by electric and mag-netic brain responses. Nature, 385, 432–4.

Naatanen, R., & Winkler, I. (1999). The concept of auditory stimulus representationin cognitive neuroscience. Psychological Bulletin, 12, 826–59.

Neininger, B., & Pulvermuller, F. (2001). The right hemisphere’s role in action wordprocessing: a double case study. Neurocase, 7 , 303–17.

Neininger, B., & Pulvermuller, E. (2002). Word category–specific deficits after lesionsin the right hemisphere. Neuropsychologia, in press.

Neville, H., Nicol, J. L., Barss, A., Forster, K. I., & Garrett, M. F. (1991). Syntacticallybased sentence processing classes: evidence from event-related brain potentials. Journal of Cognitive Neuroscience, 3, 151–65.

Neville, H. J., Bavelier, D., Corina, D., Rauschecker, J., Karni, A., Lalwani, A., Braun,A., Clark, V., Jezzard, P., & Turner, R. (1998). Cerebral organization for languagein deaf and hearing subjects: biological constraints and effects of experience.Proceedings of the National Academy of Sciences, USA, 95, 922–9.

Neville, H. J., Mills, D. L., & Lawson, D. S. (1992). Fractionating language: differentneural subsystems with different sensitive periods. Cerebral Cortex, 2, 244–58.

Newman, A. J., Bavelier, D., Corina, D., Jezzard, P., & Neville, H. J. (2002). A criticalperiod for right hemisphere recruitment in American Sign Language processing.Neuroscience, 5, 76–80.

Nobre, A. C., & McCarthy, G. (1994). Language-related EPRs: scalp distributions andmodulationbywordtypeandsemanticpriming. Journal of Cognitive Neuroscience,

6, 233–255.Noppeney, U., & Price, C. J. (2002). Retrieval of visual, auditory, and abstract seman-

tics. Neuroimage, 15, 917–26.Norris, D., McQueen, J. M., & Cutler, A. (2000). Merging information in speech

recognition: feedback is never necessary. Behavioral and Brain Sciences, 23,299–370.

Osterhout, L. (1997). On the brain response to syntactic anomalies: manipulations of 

word position and word class reveal individual differences. Brain and Language, 59(3), 494–522.

Osterhout, L., & Holcomb, P. J. (1992). Event-related brain potentials elicited bysyntactic anomaly. Journal of Memory and Language, 31, 785–806.

Page, M. P., & Norris, D. (1998). The primacy model: a new model of immediate serialrecall. Psychological Review, 105, 761–81.

Palm, G. (1980). On associative memory. Biology and Cybernetics, 36(1), 19–31.

Page 306: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 306/331

References 289

Palm, G. (1981). Towards a theory of cell assemblies.Biology and Cybernetics, 39(3),181–94.

Palm, G. (1982). Neural assemblies. Berlin: Springer.

Palm, G. (1990). Local learning rules and sparse coding in neural networks. InR. Eckmiller (Ed.), Advanced neural computers (pp. 145–50). Amsterdam:Elsevier.

Palm, G. (1993a). Cell assemblies, coherence, and corticohippocampal interplay.Hippocampus, 3(Spec No), 219–25.

Palm, G. (1993b). On the internal structure of cell assemblies. In A. Aertsen (Ed.),Brain theory: spatio-temporal aspects of brain function (pp. 261–70). Amsterdam:Elsevier.

Palm, G., & Sommer, F. T. (1995). Associative data storage and retrieval in neural

networks. In E. Domany & J. L. van Hemmen & K. Schulten (Eds.), Models of neural networks III (pp. 79–118). New York: Springer Verlag.

Pandya, D. N., & Yeterian, E. H. (1985). Architecture and connections of corticalassociation areas. In A. Peters & E. G. Jones (Eds.), Cerebral cortex, Vol. 4. Asso-

ciation and auditory cortices (pp. 3–61). London: Plenum Press.Patterson, K., & Hodges, J. R. (2001). Semantic dementia. In J. L. McClelland (Ed.),

The international encyclopaedia of the social and behavioural sciences. Section:

disorders of the adult brain. New York: Pergamon Press.Paus, T., Perry, D. W., Zatorre, R. J., Worsley, K. J., & Evans, A. C. (1996). Modulation

of cerebral blood flow in the human auditory cortex during speech: role of motor-to-sensory discharges. European Journal of Neuroscience, 8, 2236–46.

Penfield, W., & Rassmussen, T. (1950). The cerebral cortex of man. New York:Macmillan.

Penfield, W., & Roberts, L. (1959). Speech and brain mechanisms. Princeton, NJ:Princeton University Press.

Perani, D., Schnur, T., Tettamanti, M., Gorno-Tempini, M., Cappa, S. F., & Fazio, F.(1999). Word and picture matching: a PET study of semantic category effects.Neuropsychologia, 37 , 293–306.

Petersen, S., & Fiez, J. A. (1993). The processing of single words studied with positronemission tomography. Annual Review in Neuroscience, 16, 509–30.

Petersen, S., Fox, P., Posner, M., Mintun, M., & Raichle, M. (1989). Positron emis-sion tomography studies of the processing of single words. Journal of Cognitive

Neuroscience, 1, 153–70.Petitto, L. A., Zatorre, R. J., Gauna, K., Nikelski, E. J., Dostie, D., & Evans, A. C.

(2000). Speech-like cerebral activity in profoundly deaf people processing signedlanguages: implications for the neural basis of human language.Proceedings of the

National Academy of Sciences, USA, 97 , 13961–6.

Petri, C. A. (1970). Kommunikation mit Automaten. Dissertation: Universitat Bonn.Pick, A. (1913). Die agrammatischen Sprachst   orungen. Studien zur psychologischen

Grundlegung der Aphasielehre. Berlin: Springer.Pickering, M. J., & Branigan, H. P. (1999). Syntactic priming in language production.

Trends in Cognitive Sciences, 3, 136–41.Pinker, S. (1994). The language instinct. How the mind creates language. New York:

Harper Collins Publishers.

Page 307: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 307/331

290 References

Pinker, S. (1997). Words and rules in the human brain. Nature, 387 , 547–8.Plaut, D. C., & Shallice, T. (1993). Deep dyslexia: a case study of connectionist neu-

ropsychology. Cognitive Neuropsychology, 10, 377–500.

Poizner, H., Bellugi, U., & Klima, E. S. (1990). Biological foundations of language:Clues from sign language. Annual Review in Neuroscience, 13, 283–307.

Posner, M. I., & DiGirolamo, G. J. (1999). Flexible neural circuitry in word processing.Behavioral and Brain Sciences, 22, 299–300.

Preissl, H., Pulvermuller, F., Lutzenberger, W., & Birbaumer, N. (1995). Evokedpotentials distinguish nouns from verbs. Neuroscience Letters, 197 , 81–3.

Previc, F. H. (1991). A general theory concerning the prenatal origins of cerebrallateralization in humans. Psychological Review, 98, 299–334.

Price, C. J. (2000). The anatomy of language: contributions from functional neu-

roimaging. Journal of Anatomy, 197 Pt 3, 335–59.Price, C. J., Warburton, E. A., Moore, C. J., Frackowiak, R. S., & Friston, K. J. (2001).

Dynamic diaschisis: anatomically remote and context-sensitive human brain le-sions. Journal of Cognitive Neuroscience, 13, 419–29.

Price, C. J., Wise, R. J. S., & Frackowiak, R. S. J. (1996). Demonstrating the im-plicit processing of visually presented words and pseudowords.Cerebral Cortex, 6,62–70.

Pulvermuller, F. (1992). Constituents of a neurological theory of language.Concepts

in Neuroscience, 3, 157–200.

Pulvermuller, F. (1993). On connecting syntax and the brain. In A. Aertsen (Ed.),Brain theory – spatio-temporal aspects of brain function (pp. 131–45). New York:Elsevier.

Pulvermuller, F. (1994). Syntax und Hirnmechanismen. Perspektiven einer multi-disziplinaren Sprachwissenschaft. Kognitionswissenschaft, 4, 17–31.

Pulvermuller, F. (1995). Agrammatism: behavioral description and neurobiologicalexplanation. Journal of Cognitive Neuroscience, 7 , 165–81.

Pulvermuller, F. (1998). On the matter of rules. Past tense formation as a test-casefor brain models of language. Network: Computation in Neural Systems, 9, R1–51.

Pulvermuller, F. (1999a). Lexical access as a brain mechanism.Behavioral and BrainSciences, 22, 50–2.

Pulvermuller, F. (1999b). Words in the brain’s language. Behavioral and Brain

Sciences, 22, 253–336.Pulvermuller, F. (2000). Syntactic circuits: how does the brain create serial order in

sentences? Brain and Language, 71(1), 194–9.Pulvermuller, F. (2001). Brain reflections of words and their meaning. Trends in

Cognitive Sciences, 5, 517–24.Pulvermuller, F. (2002). A brain perspective on language mechanisms: from discrete

neuronal ensembles to serial order. Progress in Neurobiology, 67 , 85–111.Pulvermuller, F., Assadollahi, R., & Elbert, T. (2001). Neuromagnetic evidence for

early semantic access in word recognition. European Journal of Neuroscience,

13(1), 201–5.Pulvermuller, F., Birbaumer, N., Lutzenberger, W., & Mohr, B. (1997). High-

frequency brain activity: its possible role in attention, perception and languageprocessing. Progress in Neurobiology, 52, 427–45.

Pulvermuller, F., Eulitz, C., Pantev, C., Mohr, B., Feige, B., Lutzenberger, W.,

Page 308: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 308/331

References 291

Elbert, T., & Birbaumer, N. (1996). High-frequency cortical responses reflect lex-ical processing: an MEG study. Electroencephalography and Clinical Neurophysi-

ology, 98, 76–85.

Pulvermuller, F., Harle, M., & Hummel, F. (2000). Neurophysiological distinction of verb categories. Neuroreport, 11(12), 2789–93.

Pulvermuller, F., Hummel, F., & Harle, M. (2001). Walking or Talking?: Behavioraland neurophysiological correlates of action verb processing. Brain and Language,

78, 143–68.Pulvermuller, F., Keil, A., & Elbert, T. (1999). High-frequency brain activity: percep-

tion or active memory? Trends in Cognitive Sciences, 3, 250–2.Pulvermuller, F., Kujala, T., Shtyrov, Y., Simola. J., Tiitinen, H., Alku, P., Alho, K.,

Martinkauppi, S., Ilmoniemi, R. J., & Naatanen, R. (2001). Memory traces for

words as revealed by the mismatch negativity. NeuroImage, 14, 607–16.Pulvermuller, F., Lutzenberger, W., & Preissl, H. (1999). Nouns and verbs in the

intact brain: evidence from event-related potentials and high-frequency corticalresponses. Cerebral Cortex, 9, 498–508.

Pulvermuller, F., & Mohr, B. (1996). The concept of transcortical cell assemblies: a keyto the understanding of cortical lateralization and interhemispheric interaction.Neuroscience and Biobehavioral Reviews, 20, 557–66.

Pulvermuller, F., Mohr, B., & Schleichert, H. (1999). Semantic or lexico-syntacticfactors: what determines word-class specific activity in the human brain?

Neuroscience Letters, 275, 81–4.Pulvermuller, F., Mohr, B., Sedat, N., Hadler, B., & Rayman, J. (1996). Word class

specific deficits in Wernicke’s aphasia. Neurocase, 2, 203–12.Pulvermuller, F., & Preissl, H. (1991). A cell assembly model of language. Network:

Computation in Neural Systems, 2, 455–68.Pulvermuller, F., Preissl, H., Lutzenberger, W., & Birbaumer, N. (1996). Brain

rhythms of language: nouns versus verbs. European Journal of Neuroscience, 8,937–41.

Pulvermuller, F., & Schumann, J. H. (1994). Neurobiological mechanisms of language

acquisition. Language Learning, 44, 681–734.Rastle, K., Davis, M. H., Marslen-Wilson, W. D., & Tyler, L. K. (2000). Morphological

and semantic effects in visual word recognition: a time-course study.Language and

Cognitive Processes, 15, 507–37.Rauschecker, J. P., & Singer, W. (1979). Changes in the circuitry of the kitten visual

cortex are gated by postsynaptic activity. Nature, 280, 58–60.Redlich, A. N. (1993). Redundancy reduction as a strategy for unsupervised learning.

Neural Computation, 3, 289–304.Reichardt, W., & Varju, D. (1959). Ubertragungseigenschaften im Auswertesystem

fur das Bewegungssehen. Zeitschrift f   ur Naturforschung, 14b, 674–89.Rizzolatti, G., & Arbib, M. A. (1998). Language within our grasp. Trends in Neuro-

 sciences, 21, 188–94.Rizzolatti, G., Fadiga, L., Gallese, V., & Fogassi, L. (1996). Premotor cortex and the

recognition of motor actions. Cognitive Brain Research, 3, 131–141.Rizzolatti, G., Luppino, G., & Matelli, M. (1998). The organization of the cortical mo-

tor system: new concepts. Electroencephalography and Clinical Neurophysiology,

106, 283–96.

Page 309: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 309/331

292 References

Rosch, E., Mervis, C. B., Gray, W., Johnson, D., & Boyes-Bream, P. (1976). Basicobjects in natural categories. Cognitive Psychology, 8, 382–439.

Rosenblatt, F. (1959). Two theorems of statistical separability in the perceptron. In

M. Selfridge (Ed.), Mechanisation of thought processes: proceedings of a sympo- sium held at the National Physical Laboratory. London: HMSO.

Rugg, M. D. (1983). Further study of the electrophysiological correlates of lexicaldecision. Brain and Language, 19, 142–52.

Rugg, M. D. (1990). Event-related potentials dissociate repetition effects of high-and low-frequency words. Memory and Cognition, 18, 367–79.

Rumelhart, D. E., Hinton, G., & Williams, R. (1986). Learning internal representa-tions by backpropagation. In D. E. Rumelhart & J. L. McClelland (Eds.), Parallel 

distributed processing: explorations in the mircrostructure of cognition. Cambridge,

MA: MIT Press.Rumelhart, D. E., & McClelland, J. L. (1986). On learning the past tense of English

verbs. In J. L. McClelland & D. E. Rumelhart (Eds.), Parallel distributed processing:

explorations in the microstructure of cognition. Cambridge, MA: MIT Press.Rumelhart, D. E., & McClelland, J. L. (1987). Learning the past tense of English

verbs: implicit rules or parallel distributed processing. In B. MacWhinney (Ed.),Mechanisms of language acquisition. Hillsdale, NJ: Erlbaum.

Saffran, J. R., Aslin, R. N., & Newport, E. L. (1996). Statistical learning by 8-month-old infants. Science, 274(5294), 1926–8.

Salmelin, R., Helenius, P., & Kuukka, K. (1999). Only time can tell – words in context.Behavioral and Brain Sciences, 22, 300.

Scheibel, A. B., Paul, L. A., Fried, I., Forsythe, A. B., Tomiyasu, U., Wechsler, A.,Kao, A., & Slotnick, J. (1985). Dendritic organization of the anterior speech area.Experimental Neurology, 87 , 109–17.

Schnelle, H. (1996a). Approaches to computational brain theories of language – areview of recent proposals. Theoretical Linguistics, 22, 49–104.

Schnelle, H. (1996b).Die Natur der Sprache. Die Dynamik der Prozesse des Sprechens

und Verstehens (2 ed.). Berlin, New York: Walter de Gruyter.

Schnelle, H. (1996c). Net-linguistic approaches to linguistic structure, brain topogra-phy, and cerebral processes.

Schumann, J. H. (1997). The neurobiology of affect in language. Oxford: BlackwellPublishers.

Seldon, H. L. (1985). The anatomy of speech perception. Human auditory cortex. InA. Peters & E. G. Jones (Eds.), Cerebral cortex, Vol. 5. Association and auditory

cortices (pp. 273–327). London: Plenum Press.Shallice, T. (1988). From neuropsychology to mental structure. New York: Cambridge

University Press.

Shannon, C. E., & Weaver, W. (1949). The mathematical theory of communication.Urbana: University of Illinois Press.

Shastri, L., & Ajjanagadde, V. (1993). From simple associations to systematic reason-ing: a connectionist representation of rules, variables and dynamic bindings usingtemporal synchrony. Behavioral and Brain Sciences, 16, 417–94.

Shtyrov, Y., Kujala, T., Palva, S., Ilmoniemi, R. J., & Naatanen, R. (2000). Discrimi-nation of speech and of complex nonspeech sounds of different temporal structurein the left and right cerebral hemispheres. NeuroImage, 12(6), 657–63.

Page 310: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 310/331

References 293

Shtyrov, Y., & Pulvermuller, F. (2002a). Memory traces for inflectional affixesas shown by the mismatch negativity. European Journal of Neuroscience, 15,1085–91.

Shtyrov, Y., & Pulvermuller, F. (2002b). Neurophysiological evidence of memorytraces for words in the human brain. Neuroreport. In press.

Singer, W., Engel, A. K., Kreiter, A. K., Munk, M. H. J., Neuenschwander, S., &Roelfsema, P. R. (1997). Neuronal assemblies: necessity, signature and detectabi-lity. Trends in Cognitive Sciences, 1, 252–62.

Singer, W., & Gray, C. M. (1995). Visual feature integration and the temporal corre-lation hypothesis. Annual Review in Neuroscience, 18, 555–86.

Skrandies, W. (1998). Evoked potential correlates of semantic meaning – a brainmapping study. Cognitive Brain Research, 6, 173–83.

Skrandies, W. (1999). Early effects of semantic meaning on electrical brain activity.Behavioral and Brain Sciences, 22, 301.

Sommer, F. T., & Wennekers, T. (2000). Modelling studies on the computationalfunction of fast temporal structure in cortical circuit activity. Journal of Physiology,

Paris, 94, 473–88.Sougn´ e, J. (1998). Connectionism and the problem of multiple instantiation. Trends

in Cognitive Sciences, 2, 183–9.Spitzer, M., Kischka, U., Guckel, F., Bellemann, M. E., Kammer, T., Seyyedi,

S., Weisbrod, M., Schwartz, A., & Brix, G. (1998). Functional magnetic reso-

nance imaging of category-specific cortical activation: evidence for semantic maps.Cognitive Brain Research, 6, 309–19.

Steinmetz, H., Volkmann, J., Jancke, L., & Freund, H. J. (1991). Anatomical left–right asymmetry of language-related temporal cortex is different in left- and right-handers. Annals of Neurology, 29, 315–19.

Sterr, A., Muller, M. M., Elbert, T., Rockstroh, B., Pantev, C., & Taub, E. (1998).Perceptual correlates of changes in cortical representation of fingers in blindmultifinger Braille readers. Journal of Neuroscience, 18(11), 4417–23.

Swinney, D., Onifer, W., Prather, P., & Hirshkowitz, M. (1979). Semantic fascilation

across sensory modalities in the processing of individual words and sentences.Memory and Cognition, 7 , 159–65.

Tallon-Baudry, C., & Bertrand, O. (1999). Oscillatory gamma activity in humans andits role in object representation. Trends in Cognitive Sciences, 3, 151–61.

Tesniere, L. (1953). Esquisse d’une syntax structurale. Paris: Klincksieck.Tesniere, L. (1959). El   ements de syntaxe structurale. Paris: Klincksieck.Tranel, D., & Damasio, A. R. (1999). The neurobiology of knowledge retrieval.

Behavioral and Brain Sciences, 22, 303.Tsumoto, T. (1992). Long-term potentiation and long-term depression in the neocor-

tex. Progress in Neurobiology, 39, 209–28.Ullman, M., Corkin, S., Coppola, M., Hickok, G., Growdon, J., Koroshetz, W., &

Pinker, S. (1997). A neural dissociation within language: evidence that the mentaldictionary is part of declarative memory, and that grammatical rules are processedby the procedural system. Journal of Cognitive Neuroscience, 9, 266–76.

Vaadia, E., Haalman, I., Abeles, M., Bergman, H., Prut, Y., Slovin, H., & Aertsen, A.(1995). Dynamics of neuronal interactions in monkey cortex in relation tobehavioural events. Nature, 373, 515–18.

Page 311: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 311/331

294 References

Varju, D., & Reichardt, W. (1967). Ubertragungseigenschaften im Auswertesystemfur das Bewegungssehen II. Zeitschrift f   ur Naturforschung, 22b, 1343–51.

von der Malsburg, C. (1985). Nervous structures with dynamical links. Bericht der 

Bunsengesellschaft f   ur Physikalische Chemie, 89, 703–10.von der Malsburg, C. (1995). Binding in models of perception and brain function.

Current Opinions in Neurobiology, 5, 520–26.Waibel, A. H., Hanazawa, T., Hinton, G., Shikano, K., & Lang, K. J. (1995). Phoneme

recognition using time-delay neural networks. In Y. Chauvin & D. E. Rumelhart(Eds.), Backpropagation: theory, architectures, and applications. Developments in

connectionist theory (pp. 35–61). Hillsdale, NJ: Lawrence Erlbaum Associates.Warburton, E., Wise, R. J. S., Price, C. J., Weiller, C., Hadar, U., Ramsay, S., &

Frackowiak, R. S. J. (1996). Noun and verb retrieval by normal subjects: studies

with PET. Brain, 119, 159–79.Warrington, E. K., & McCarthy, R. A. (1983). Category specific access dysphasia.

Brain, 106, 859–78.Warrington, E. K., & McCarthy, R. A. (1987). Categories of knowledge: further

fractionations and an attempted integration. Brain, 110, 1273–96.Warrington, E. K., & Shallice, T. (1984). Category specific semantic impairments.

Brain, 107 , 829–54.Weiller, C., Isensee, C., Rijntjes, M., Huber, W., Muller, S., Bier, D., Dutschka, K.,

Woods, R. P., North, J., & Diener, H. C. (1995). Recovery from Wernicke’s aphasia:

a positron emission tomography study. Annals of Neurology, 37 , 723–32.Wennekers, T., & Palm, G. (1996). Controlling the speed of synfire chains.Proceedings

of the International Conference on Artificial Neural Networks, 451–6.Wermter, S., & Elshaw, M. (2002). A neurocognitively inspired modular approach to

self-organisation of action verb processing. Connection Science, In press.Wernicke, C. (1874). Der aphasische Symptomencomplex. Eine psychologische Studie

auf anatomischer Basis. Breslau: Kohn und Weigert.Weyerts, H., Penke, M., Dohrn, U., Clahsen, H., & Munte, T. (1997). Brain potentials

indicate differences between regular and irregular German plurals.NeuroReport,

8, 957–62.Wickelgren, W. A. (1969). Context-sensitive coding, associative memory, and serial

order in (speech) behavior. Psychological Review, 76, 1–15.Wickens, J. R. (1993). A theory of the striatum. Oxford: Pergamon Press.Willshaw, D. J. Buneman, O. P., & Longuet-Higgins, H. C. (1969). Non-holographic

associative memory. Nature, 222 (197), 960–2.Willwacher, G. (1976). Faehigkeiten eines assoziativen Speichersystems im Vergleich

zu Gehirnfunktionen. Biological Cybernetics, 24, 181–98.Willwacher, G. (1982). Storage of a temporal pattern sequence in a network.Biologi-

cal Cybernetics, 43, 115–26.Winograd, T. (1983). Language as a cognitive process, vol. 1: Syntax. Reading, MA:

Addision-Wesley.Wise, R., Chollet, F., Hadar, U., Fiston, K., Hoffner, E., & Frackowiak, R. (1991).

Distributionof cortical neural networks involved in word comprehension andwordretrieval. Brain, 114, 1803–17.

Wittgenstein, L. (1953). Philosophical Investigations. Oxford: Blackwell Publishers.

Page 312: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 312/331

References 295

Woods, B. T. (1983). Is the left hemisphere specialized for language at birth? Trends

in Neurosciences, 6, 115–17.Woods, W. A. (1973). An experimental parsing system for transition network gram-

mars. In R. Rustin (Ed.), Natural language processing (pp. 111–154). New York:Algorithmics Press.

Young, M. P., Scannell, J. W., & Burns, G. (1995). The analysis of cortical connectivity.

Heidelberg: Springer.Zaidel, E. (1976). Auditory vocabulary of the right hemisphere following brain

bisection or hemideocortication. Cortex, 12, 191–211.Zaidel, E. (1985). Language in the right hemisphere. In D. F. Benson & E. Zaidel

(Eds.), The dual brain (pp. 205–31). New York: Guilford.Zaidel, E. Kasher, A., Soroker, N., Batori, G., Giora, R., & Graves, D. (2000). Hemi-

spheric contributions to pragmatics. Brain Cognition, 43(1–3), 438–43.Zaidel, E., & Rayman, J. (1994). Interhemispheric control in the normal brain: evi-

dence from redundant bilateral presentation. In C. Umilta & M. Moscovitch (Eds.), Attention and performance XV: concious and subconcious information processing

(pp. 477–504). Boston, MA: MIT Press.Zatorre, R. J., Evans, A. C., Meyer, E., & Gjedde, A. (1992). Lateralization of pho-

netic and pitch discremination in speech processing. Science, 256, 846–49.Zhou, X., & Marslen-Wilson, W. (2000). The relative time course of semantic and

phonological activation in reading Chinese. Journal of Experimental Psychology:

Learning, Memory and Cognition, 26(5), 1245–65.Zhou, Y. D., & Fuster, J. M. (2000). Visuo-tactile cross-modal associations in corti-

cal somatosensory cells. Proceedings of the National Academy of Sciences, USA,

97 (17), 9777–82.Zipser, D., Kehoe, B., Littlewort, G., & Fuster, J. (1993). A spiking network model

of short-term active memory. Journal of Neuroscience, 13(8), 3406–20.Zohary, E. (1992). Population coding of visual stimuli by cortical neurons tuned to

more than one dimension. Biological Cybernetics, 66, 265–72.

Page 313: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 313/331

Page 314: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 314/331

 Abbreviations

General

“is rewritten as”; symbol for formulating rules of phrasestructure grammars

→→ “activates”; symbol for asymmetric reciprocal connection,

the basic element for establishing serial order in syntactic

circuits

↔ “activates”; symbol for symmetric reciprocal connection

⇒ “is followed by,” “causes”

A, B, . . . words or morphemes

a, b, . . . lexical categories

α, β . . . neuronal elements – neurons, cell assemblies, neuronal sets,or sequence sets

AB sequence of symbols A and B

ab sequence of symbols belonging to categories a and b

α pi i th sequence set sensitive to information in the past (“past

set”) representing sequences of symbols that end with a

symbol from category a

α f j  j th sequence set sensitive to information in the future(“future set”) representing sequences of symbols that begin

with a symbol from category a

S p, Sq, . . . Neuronal sets (input or sequencing units)

297

Page 315: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 315/331

298 Abbreviations

Fragments of Neuronal Grammar

α ↔ λ assignment formula: the word/morpheme representation α

is assigned to the lexical category representation λ

α( p1, . . . , pm/*/ f 1, . . . , f n)

valence formula: lexical category representation α includesm backward-looking past sets p1, . . . , pm and n forward-

looking future sets f 1, . . . , f n.

α( f i)→→β( p j ) there is a directed connection from the forward-looking

sequence set f i belonging to category representation α

to the backward-looking sequence set p j  of category

representation β.

Syntactic and Lexical Categories

Note: Abbreviations of lexical categories are used to refer to sets of lexical

items and for labeling their neuronal representations.

S sentence

NP noun phrase or nominal phrase

VP verb phrase or verbal phrase

Det determiner or article

N noun

Prop proper name

N1, Nnom nominative noun

N3, Ndat dative noun

N4, Nakk accusative noun

V verb

Vi intransitive verbVt transitive verb

V(t)sub transitive verb in subordinate sentence

Vip intransitive particle verb

Page 316: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 316/331

Abbreviations 299

Vtp transitive particle verb (tp is sometimes omitted)

V1, V14, V134,

etc.

verb subcategories specified in terms of the noun comple-

ments required (V14 = verb requiring N1 and N4)

Vp verb particle

Vs verb suffix

Tr translative

Page 317: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 317/331

Page 318: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 318/331

 Author Index 

Abeles, 148–150, 157, 178

Aboitiz, 54

Aertsen, 148

Ahissar, 19

Ajjanagadde, 145

Ajukiewicz, 214, 237

Alho, 16

Angrilli, 65

Arbib, 27

Aslin, 51

Assadollahi, xiii, 61, 63, 65

Bach, 129, 133

Bak, 39, 59

Baker, 207

Balota, 45

Bar-Hillel, 209

Barlow, 23–24, 112, 160–161

Barss, 267

Bates, 207

Baudry, 157

Bavelier, 38

Bellugi, 38

Benson, 116

Betrand, 32, 53, 157,

274

Bienenstock, 241

Biermann, 53

Birbaumer, 32, 53, 60

Bird, 39

Bock, 165

Bodner, 27

Bogen, xiii

Borst, 177

Braitenberg, xii, 1, 9, 11–13, 16, 20–21,

25, 37, 54–55, 64, 76, 78–79, 90, 105,

151, 157, 160, 240, 246

Branigan, 165

Brent, 51

Bressler, 157, 274

Broca, 13, 28, 34–35, 39

Brodmann, 13–14, 35, 47

Brown, 60, 76, 118, 167, 267

Bruwer, 166

Buonomano, 16, 38, 166

Burger-Judisch, 60

Burns, 17

Caillieux, xiii

Caplan, 33, 167

Cappa, 59

Caramazza, 37, 59

Cartwright, 51

Changeux, 166

Chao, 60

Charniak, 106

Cheour, 51

Chiarello, 43

Chomsky, 106, 124–125, 128–129, 137, 245,

271–273

Corina, 38

Crespi-Reghizzi, xii

Creutzfeldt, 53

Cruse, 166

Cutler, 110

Dale, 53, 55

Damasio, 24, 39, 46–47, 60

301

Page 319: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 319/331

302 Author Index

Daniele, 39, 40, 59

Daugherty, 114

De Agostini, 40

De Renzi, 28, 34De Saussure, 270

Deacon, 18

DeFelipe, 12

Dehaene, 41, 166

Dehaene-Lambertz, 41

Dell, 108–110

Devlin, 60

Diesch, 53

Diesmann, 148

DiGirolamo, 46–47, 90Dilthey, 199

Dobel, 40

Dum, 15

Dumais, 57

Dungelhoff, 83

Egelhaaf, 177

Elbert, xiii, 16, 61, 274

Ellis, 92–94, 108

Elman, 114, 115, 166, 207Elshaw, 73

Epstein, 46–47

Eulitz, 53

Evett, 83

Farah, 119

Farinas, 12

Federmeier, 60–61

Fiez, 40, 44, 45

Forde, 39, 60, 119Forster, 267

Frackowiak, 53

Freeman, 274

Freud, 1, 48

Friederici, 167, 267

Fritz, xiii

Fry, 37

Fuster, xiii, 19, 27–28, 30–31, 55, 57, 64, 79,

163, 170

Gaifman, 139, 214

Galuske, 18, 41

Garrett, 108, 267

Gazdar, 137, 144

Geschwind, 41, 46–47, 115

Gewaltig, 148

Gill, 60

Glaser, 83

Golinkoff, 60

Goodglass, 40

Goulet, 40Grabowski, 60

Gray, 274

Groothusen, 167, 267

Gunter, 267

Hacker, 207

Hadler, 36

Haegeman, 125, 127, 237

Hagoort, 53, 118, 167, 267

Hahne, 167, 267Hannequin, 40

Hare, 114, 120

Harle, 62

Harris, 51, 124, 135, 137, 155

Hasbrooke, 43

Hauk, xiii, 62

Hawkins, xiii

Hayes, 41

Hays, 139, 214, 237

He, 15Hebb, 19, 22, 23, 75–76, 156, 169

Hecaen, 40

Heck, xiii, 160

Heim, 79

Helenius, 46

Hellwig, 16

Helmholtz, 275

Heringer, 139, 143, 214, 246

Hetherington, 237

Hill, 161Hinton, 114, 119

Hirsch-Pasek, 60

Hodges, 59–60, 116

Holcomb, 167, 267

Hubel, 77, 160

Hummel, 62

Humphreys, 39, 60, 83, 119

Illmoniemi, xiii

Indefrey, 167

Jackendoff, 127

Jackson, 33, 39–40

Jacobs, 18, 41

Joanette, 40

Johnson, 207

Joshi, 137

Page 320: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 320/331

Author Index 303

Kandel, 19–20

Kaplan, 133

Karmiloff-Smith, 207

Keil, 274Keyser, 34

Kiefer, 59, 61, 64

King, 269

Kleene, 97, 100, 114, 159, 186

Klein, 137

Kleinfeld, 166

Klima, 38

Kluender, 269, 273

Koenig, 60

Kolk, 34Korpilahti, 54–55

Krause, 53

Kreiter, 32

Kroger, 27

Kujala, 16, 54–55, 63

Kutas, 269, 273

Kuukka, 46

Lambek, 214

Landauer, 57Lashley, 23, 151

Lawson, 118

Le Clec’H, 65

Lee, 41

Lehmann, 60, 268

Lettich, 53

Levelt, 83, 110, 128, 131

Levick, 160–161

Levitsky, 41

Levy, 138Lewis, 41

Lichtheim, 28, 34–36, 44, 48

Locke, 51

Loebell, 165

Luce, 53

Luppino, 27

Lutzenberger, xii, 32, 53, 57, 60, 118,

158, 274

Marcel, 116

Marcus, 141

Marie, 33

Markov, 106

Marshall, 272

Marslen-Wilson, xiii, 55, 64, 86, 121, 129, 154

Martin, 60

Mason, 16

Matelli, 27

McCarthy, 39, 57, 60, 118

McClelland, 108, 119–120

McCulloch, 97, 100, 103, 107, 124, 159, 166,173, 186, 207

McQueen, 110

Mecklinger, 267

Merzenich, 16, 38

Mesulam, 1

Meyer, 83, 110

Miceli, 59

Miller, xii, 41, 79–80, 106, 179

Mills, 118

Milner, 23, 25, 53, 75, 79, 170, 274Minsky, 97,113

Mittelstadt, 43

Mohr, 32, 36, 43, 53, 61, 83

Molfese, 41, 60

Montoya, 91

Monzon-Montes, 40

Moore, 60

Morey, 165, 167

Moro, 167

Morton, 93Muller, 269

Mummery, 60, 65

Naatanen, xiii, 16, 40, 51, 54

Nadal, 166

Neininger, xiii, 39–40

Neville, 38, 118, 167, 267

Newman, 38

Newport, 51

Nicol, 267Nicoll, 16,

Nobre, 118

Nocentini, 59

Noppeney, 60

Norris, xiii, 110, 154

Ojemann, 53

Older, 86

Osterhout, xiii, 167, 267

Page, 154

Palm, xiii, 21, 25, 75, 77, 79, 90, 112

Pandya, 17–18, 110, 178

Papert, 113

Parisi, 207

Patterson, 59– 60, 116

Paus, 44

Page 321: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 321/331

304 Author Index

Penfield, 15, 62

Perani, 60

Perles, 207

Petersen, 40, 44– 45Petitto, 38–39

Petri, 133

Pfeifer, 167

Pick, 115

Pickering, 165

Pinker, 23, 120, 207

Pitts, 97, 100, 103, 107, 124, 159, 166, 173,

186, 207

Plaut, 73, 114, 119

Plunkett, 207Poizner, 38

Posner, 46–47, 90

Preissl, 53, 57, 60, 69–70, 72–73, 118

Previc, 41

Price, 40, 44, 53, 56, 60

Pullum, 137

Pulvermuller, 1, 32, 36, 39–43, 45– 47, 51–65,

69–73, 83, 91, 116, 118, 122, 129, 144,

146, 151, 158, 176, 208–209, 244,

268, 274Pushkin, 106

Quadfasel, 40

Raichle, 45

Rassmussen, 62

Rastle, 86

Rauschecker, 76

Rayman, 36, 43

Redlich, 51Reichardt, 106, 160, 162, 166, 177

Rizzolatti, 27

Roberts, 15

Rochstroh, xiii

Roelofs, 110

Rosch, 89

Rosenblatt, 113–114

Rugg, 53–54, 63, 65

Rumelhart, 108, 114, 120

Saffran, 51

Sag, 137

Salmelin, 46–47

Scannell, 17

Scheibel, xiii, 18, 41

Schleichert, 61

Schnelle, xiii, 1, 91, 133, 147, 255, 271

Schriefers, 267

Schumann, xiii, 91

Schuz, xii, 11–13, 16, 20–21, 157Schvaneveldt, 83

Sedat, 36, 116

Seldon, 41

Shallice, 23, 39, 60, 66, 73, 114, 119

Shamir, 209

Shannon, 51, 106

Shapiro, 237

Shastri, 145

Shtyrov, xiii, 12–13, 40, 54–55, 118

Silveri, 59Singer, 32, 76, 274

Skrandies, 47, 63, 268

Sommer, 25, 73

Sompolinsky, 166

Sougne, 145

Spitzer, 59– 60

Steinmetz, 41

Sterr, 16

Stratford, 16

Strick, 15Sultan, 160

Swinney, 84

Tallal, 45

Tallon-Baudry, 32, 53, 157, 274

Taylor, 83

Ter Keurs, 118

Tesniere, 124, 139, 214, 237, 246

Tranel, 39, 46

Tsumoto, 19–20, 76, 237Tyler, 55, 64, 86, 121

Ullman, 121

Vaadia, 148

Van Grunsven, 34

Varju, 160, 162, 166

Vignolo, 28, 34

Villa, 59

Von der Malsburg, 53, 274

Waibel, 114

Warburton, 60

Warren, 154

Warrington, 39, 57, 60, 73

Weaver, 51, 106

Page 322: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 322/331

Author Index 305

Weiller, 40

Wennekers, 73

Wermter, 73

Wernicke, 28, 34–36Weyerts, 267

Wickelgren, 151

Wickens, 79–80

Williams, 114

Willwacher, 166

Winkler, 54, 79

Winograd, 107, 133

Wise, 45, 53

Wittgenstein, 26, 89–90, 154

Woods, 40, 133

Yeterian, 17–18, 110, 178Young, 17, 92–94, 108, 110, 178

Zaidel, 42–43

Zatorre, 40, 44

Zhou, 27

Zipser, 163

Zohary, 23

Zurif, 37

Page 323: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 323/331

Page 324: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 324/331

Subject Index 

A

abstraction, 233

action, 27, 28, 60, 63

associations, 58–59

potential, 18, 19, 32

related nouns, 61

related verbs, 57, 62, 63

words, 5, 56, 58, 63

activity

high frequency, see also gamma band,

responses, 52–56, 61, 157–158, 169, 274

states, 169, 170, 173

vectors, 77, 112

waves, 220

adjectives, 104, 116

adjuncts, 127, 140, 235, 236, 237, 238, 245,

247, 249, 262

affixes, 65, 189

inflectional, 115, 203

agrammatism, 73, 115, 118

agreement, 7, 144, 164, 225, 231, 249

subject-verb, 248

algorithms, 96, 186, 190, 207, 209, 248,

262, 271

symbolic, 119

syntactic, 188, 208

ALL, logical quantor, 104, 132

allophones, 152, 153

Alzheimer’s disease, 121

ambiguity, 187, 250, 192, 196, 197, 250, 253,

254, 61

amygdala, 91

AND, logical operation, 98, 99, 103

Angular gyrus, 38, 47

animal names, 57–58

anomia, 34, 73, 116, 118

anti-Hebb learning, 76

aphasia, 1, 2, 13, 34, 115, 4, 28, 33, 36,

40, 72

amnesic, anomia, 34

Broca’s, 34, 121

conduction, 36

mixed transcortical, 36

motor, 34

multimodal character of, 34

total, 34

Wernicke, 34–37

aphasic patients, see also aphasia, 66, 110

areas, see also language areas, perisylvian

areas, temporal areas

Broca’s, 35, 36, 38, 45, 72, 118

frontal, 39, 45, 80

nonprimary, 22, 24

occipital, 44–47, 91

parietal, 35

prefrontal, 27, 59, 69

premotor, 60, 69

primary, 15–16, 18, 20, 22, 72, 111

relay, 21

somatosensory, 17

visual, 58, 59, 60, 91

Wernicke’s, 44–45, 70–72

arousal, 39

article, see also determiner, 117, 115, 126

articulations, 69, 37, 154

articulators, 37, 108

articulatory patterns, 35

assignment formulas, 209, 210, 217, 226, 227,

250, 256

astonomer, 270

307

Page 325: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 325/331

308 Subject Index

A-system or cortical connections, 11, 12,

13, 16

attention, 54, 56

automata theory, 107axon, 10, 11, 12, 18, 19, 41, 98, 193

axonal conduction delays, 54

B

babbling, 37, 51

basal dendrites, 12, 16

bilateral advantage, 43

binding, 7, 24, 70, 112

brain

function, 96lesions, 48, 67, 114

lesions, focal, 23, 69–70

mechanisms, 147, 245, 272, 274, 275

theory, 96, 274

Broca, 36, 37, 39, 69, 72

aphasia, 34, 121

area, 35, 36, 38, 45, 72, 118

C

cardinal cells, 24, 98, 100, 102, 112, 113,159, 160

neuron, 99, 103

categorical grammar, 214

categories, of concepts/words, 39, 43, 49, 57,

60, 65, 88, 119

categorization, 161

category

basic level, 89

conceptual, 57, 60

differences, 65lexical, see lexical categories

specific brain processes, 57–65

syntactic, 125–131, 138–144,

143–145

cell

assemblies, see also neuronal ensemble,

22–26, 49, 62–64, 72–77, 90, 97, 111,

112, 116, 121, 118, 156, 157, 168,

170, 171

body, 10, 98center embedding, 5, 7, 8, 115, 128–131, 144,

245–246, 249, 255, 262, 264

circuits, 81, 98, 99, 107, 100

syntactic, 204, 209, 227, 238

clinical tests, 39

clusters of neurons, local, 17, 21, 23

coarticulation, 154

cognitive

neuroscience, 2, 23, 73, 154, 168

processing, 23, 22, 27, 168

psychology, 3science, 24, 114

coherence, 32

coherent oscillatory brain activity, see also

high-frequency activity, 32, 220

column, 17

complement, 127, 130–134, 139, 155, 187,

188, 192, 210, 214, 232, 235, 236, 238,

245, 247, 257

complex

classification, 122event, see also events, 99, 102

comprehension, 37, 38, 59, 69, 71, 107, 108

concepts, 35, 119, 237, 241

conditional probability, 106

connectionist

models, symbolic, 97, 107–110

models, distributed, 112, 119

theories, 5, 34

constitute, 233

content word, 73, 87, 116–117context, 55, 59, 65, 84–85, 148, 150, 194, 208,

235, 266

dependence, 148, 149

dependent meanings, 90

sensitive grammar, 199

sensitive phoneme variants, 151

free grammar, 127, 128, 130, 137, 142

conversations, 245

correlation, 20, 25, 28, 57, 81, 88, 89, 207,

236, 273learning, 21, 22, 51, 66, 91, 179

cortex, 4, 9, 10, 12, 23, 37, 69–70, 72, 74–75,

77, 80–81, 83

motor, 15, 18, 37, 58, 63, 69

cortical

circuits, local, see also synfire chain, 41

laterality, see also laterality, 38

topography, 28, 56, 58, 69, 267

countability, 174

current source density analysis, 58

D

deaf, 38

deficits, category-specific, 119

definitions, 90

delayed matching to sample task, 26–27

dendrites, 10, 11, 12–13, 17–19, 98

Page 326: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 326/331

Subject Index 309

dependency, 190, 224, 231, 234, 249

crossed, 129

grammars, 5, 124, 138–143, 187, 189, 192,

211, 213–214, 246long-distance, 5, 6, 105, 136, 137–139, 155,

164, 225, 249

mutual, 82, 236

relationship, 136, 143

rules, 139, 140, 188, 191, 197

derivational affix, 86, 87

determiner, see also article, 126, 145, 188,

189, 226, 229, 231

disambiguation, 196, 199, 200

discontinuous constitiuent, 104–105,134–139, 143–144, 225, 230, 249

discrete time steps, 98

distinctive features, 152

divergence, 21

dopamine, 91

double dissociations, 5, 67, 73, 94, 118, 121,

122, 123

double stop consonants, 55

dynamics, 28–30, 32, 74, 80, 250

high frequency, 158dyslexia, deep, 116, 118

E

EEG, 32, 44, 46, 53, 54, 63

EITHER-OR, logical operation, 98–99, 100,

114, 122

electroencephalography (EEG), 40, 267

electrophysiological, 51, 53

embedding, see also center embedding, 235,

263engrams, 23, 31–33, 54, 76–77

ensembles of neurons, see cell assemblies,

epileptic seizure, 76

event related potential (ERP), 54, 58,

61, 64

events, 97, 104, 147, 165, 177, 229

excitatory, 18

existence

neuron, 104, 132–133, 137–139

proof, 96explanation, 66, 118, 122

F

family resemblance, 88, 90, 97, 154

features, see also distinctive features, 32, 85,

108

feedback control, 79

finite

automaton, 131

state device, 161–164, 174, 178–182, 184,

190, 198, 208, 218, 219state grammar, 107, 212

state automata, 97

state networks, 248

focused attention, 46

form meaning relationship, 74

frontal, 39, 45, 80

cortex, 34, 37, 46, 70, 35, 44, 50, 70–72, 118

function words, 65, 115–118, 246

functional

affix, 117attributes, 60

cortical webs, see also functional webs, 49

magnetic resonance imaging (FMR), 40, 44

units, 22, 23

webs, 4, 10, 22–30, 32, 50, 51–59, 63,

65–67, 69, 71–72, 74, 77, 81–83,

87–88, 90, 95, 111–121, 153–171,

208, 274

webs, overlapping, 84

G

gamma band responses, see also high

frequency, 52, 56, 58, 64

generalization, 155, 162

gestalt, 29

glia, 41

global, 34

grammar, 25, 129, 131, 139, 142, 154, 156,

159, 167, 174, 178, 186, 193, 224, 226,

245, 256, 257, 262algorithms, 1, 95, 107, 128, 139, 233

circuits, 5–6, 65, 141, 174, 208, 224, 234,

241, 250, 263

mechanisms, see also grammar circuits,

1, 95, 129, 163

networks, see also grammar circuits, 6, 194,

200, 215, 216, 218, 223, 233

one-sided linear, 137

regular, 93, 128–132

theories, 132, 208, 214, 237grammatical

contexts, 90

rules, see also rules, 107

string, 220

grapheme, 94

gray matter, 34

gyri, 13, 14, 23

Page 327: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 327/331

310 Subject Index

H

hemispherectomy patients, 42

hemispheres, 4, 43, 52

dominant, see also cortex, laterality, 17, 21,35, 51, 82, 118

left, 39, 40, 45, 63, 267

nondominant, 42, 43, 49

right, 43, 53

hermeneutic circle, 199

hetero-associative, see also associative

network, 75

hidden layer, 114

hidden Markov models, 107

hierarchy, 234higher cognitive processes, 22, 27

hippocampus, 19

homonymy, 87

homophony, 87

homotopic areas, 17, 21

hyperonym, 88

hyponyms, 88, 90

I

ignition threshold, 67, 69, 170, 172, 173, 182ignitions, 6, 54, 29, 32, 55, 64, 67, 69, 71, 81,

84, 90, 153, 168–184, 190, 198, 199, 201,

203–208, 210–269

imageability, 116

imaging

studies, 13, 44–48, 272

techniques, metabolic, 44, 53–56, 60,

64, 73

infants, 40, 115, 120

inflection, 121, 122, 123information, 9, 21, 24, 27–28, 34, 55–56, 64,

66, 69, 94, 101–105, 108, 111, 114,

117–119, 131–138, 145, 146, 150, 152,

165, 166, 178–181, 194, 200, 202, 207,

229, 234, 239, 246, 248, 262,

265, 266

linguistic, 110

mixing, 22

mutual, 51

inhibitory postsynaptic potentials, 18input units, 101, 218

interhemispheric interaction, 4, 43

interpretation, 199

K

kernel, 25, 240

k-limited stochastic models, 106

L

language

acquisition, 50, 192

areas, complementary language areas,49

areas, core language areas, 49, 56

areas, see also Broca area, Wernicke area,

36, 45, 48, 60

cognitive neuroscience of, see cognitive

neuroscience

comprehension, 34, 37

disturbances, see also aphasia, 33

dominance, see also laterality, 40, 116

laterality, 40production, 28, 36, 37, 44–46, 66, 108,

116

regular, 132

spoken, 33, 34, 38–39, 93

tasks, 46

theory, 8, 272

latencies, 268

laterality, 4, 41–42, 82–83

layers, 97, 100, 108, 110–112, 114, 150

learning, see also associative learning, 69, 81,88

associative, 22, 29, 50, 69, 74–75, 122, 19,

20, 56, 162, 209

of word meaning, 57, 59

rule, see also correlation, 75, 114

substitution-based associative, 162

left anterior negativity, 267

lesion

of the brain, see also brain lesions, 36,

39–42, 47, 67, 72simulations, 72

letter sound correlation, 93

letters, 93, 106, 108, 151

lexical access, 71

ambiguity, see also ambiguity, 7, 249

categories, 7, 126, 127, 139, 141, 143, 145,

146, 155, 162, 165, 168, 186–199,

203–205, 208–214, 226–229, 232, 235,

236, 244, 245, 249, 250, 254, 256, 257,

262, 264decision, 42, 58, 65

frequency, 89

nodes, 108

status, 55, 63–64

lexicon, 193

rules, 126, 135–140

Lichtheim scheme or model, 38, 39

Page 328: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 328/331

Subject Index 311

linguistic, 43

linguistic infinity position, 238

linguistics, 3, 8, 33, 124 –125, 270, 271,

275living things, 57

localist connectionist models, 145

logical

circuits, 99

computation, 208

formula, 100

operations, 97, 98, 100, 102, 107

operators, 100

logogen model, 93

long-term depression, 237

M

macaque monkeys, 17, 26–28

magnetoencephalography (MEG), 61, 54, 32,

44, 54, 64, 274

Markovian sources, 106

McCulloch-Pitts

network, 101, 107, 133, 177, 182

neurons, 97, 100, 104, 106, 108, 131,

145meaning, 48–49, 57–65, 88, 93, 109, 116–117,

121, 125, 145, 154, 188, 226, 248

referential, 93

mechanisms, 73, 79–82, 88, 124

mediated sequence processing, 159 –165, 160,

177, 181

memory, 26–31, 46, 51, 65, 75, 93, 110, 114,

129, 273

active, 30, 55, 64, 190

associative, 28–29, 50, 66, 76–79, 95, 111,112, 207, 75, 85, 171

cells, 27–30, 139, 146, 163, 170, 271

layer, 114

mechanism, 131

networks, 32

neurons, see memory cells

trace, 30, 54–56

midbrain, 91

mismatch negativity (MMN), 52–55

MMN, see mismatch negativity, 52,54–55

modeling, see neuronal networks, 3

modular models, see also modules, 73, 93,

108, 110

modular theories, see also modular models,

modules, 93–94, 107–111

modules, 66–67, 93–94, 97, 107, 207

morphemes, 23, 51, 65, 93, 104, 117, 129, 146,

147, 154, 161–165, 182, 185–188,

191–194, 205, 209, 224, 227, 228, 231,

234, 236, 238, 241, 254, 266inflectional, 267

sequences, 186

motion detectors, 162

mouse, 17

movement, 137, 257

detection, 159

multilayer networks, 100

multiple occurrences, 235, 238, 241, 262

reverberation, 7, 239–240, 245, 250, 255,

272

N

naming, 58, 65

neostriatum, 81

nerve cell, see also neuron, 9, 10

nervous system, 76, 96

network, see also neuronal network,

67, 78

associative, 120, 166

augmented transition, 107, 133dynamics, 215

memory, 75

neural

algorithms, 5, 96

assembly, 25, 81, 90

structure, 207

neuroanatomical studies, 17, 18, 41, 42

neuroanatomy, 17–18, 37–42, 247

neurobiological mechanisms, 8, 57, 91,

158neurocomputational research, 32, 165

neuroimaging, 2–4, 38, 40, 44, 50, 54

neurological model of language, 34, 48

neurology, 3, 33

neuron, 4, 9, 10–13, 17–22, 28–29, 31, 32,

50, 52, 64, 67, 71, 98, 100, 102, 104,

113

circuits, 98, 139, 207, 270, 271, 272, 224,

275

clusters, see local neuron clusters, synfirechains, 24

delay, 101–102

density, 59

group, 170, 171

mirror, 27

pyramidal, 10–13, 41, 78, 102

webs, 91

Page 329: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 329/331

312 Subject Index

neuronal

algorithms, 124

assembly, 69, 75

automaton, 208, 233dynamics, 103

ensembles, 1–4, 23, 25, 28, 51, 63,

67, 69, 75, 79, 80, 87, 90, 95, 146,

168, 171, 194, 206, 208, 247,

274

finite state automata, 245

grammar, 6 –7, 127, 137, 179, 191–196,

203–216, 224, 225, 228, 231–236,

239–251, 255, 256, 264–269

group, 25, 168mechanisms, 9, 96, 170, 247, 271

networks, 2–5, 96 –97, 115–120, 177, 207,

105, 200, 246

representations, 51, 59

representations, overlap of, 5, 74, 81–84,

86–90

sets, 168–182, 186, 187, 193, 199, 200,

201, 207, 208, 209, 215, 225, 229,

238, 240–247, 251–255, 262–265,

269, 272neurophysiology, 4, 5, 30, 32, 37, 39–40,

61–62, 147, 160, 163, 247, 265,

269

neuropsychological

data, 50, 73

dissociations, see also double dissociation,

95

imaging techniques, 44, 64, 73, 118

syndromes, 66

neuropsychology, 3, 8neuroscience, 3, 42–47, 271

nodes, 11, 109, 110

nonliving things, 57

NOT, logical operation, 98, 99

nouns, 39, 45, 58–59, 61, 60, 86, 104,

116, 126, 140, 143–144, 155, 156,

162, 187–190, 197, 198, 210, 211,

214, 226, 231, 236–238, 250–251,

253, 262

nucleus candatus, 80

O

object representation, 29, 32

OR, logical operation, 98, 99, 103

orthography, 84

oscillatory activity, see also high-frequency

activity, rhythms, 157, 274

overexcitation, 78

overgeneralizations, 120

PP600, 267

paleostriatum, 80–81

Parkinson’s disease, 121

particle, 104, 105, 106, 143, 225

verbs, 140, 225, 248, 251, 252

passive, 30, 37

past tense, 119, 120, 122, 208

perceptron, 5, 113, 114, 119, 120, 121

perisylvian

areas, 14, 35–38, 51–53, 56–58, 69–72,82–83, 118, 167

cortex, see also perisylvian areas, 51, 53,

70, 82–83, 167

lesion, 73

regions, see perisylvian areas

phonemes, 41, 51, 94, 108, 147,

151–158

phonetic judgment, 44

phonetics, 152

phonologicalprocesses, 4, 40, 151–158

webs, 56–57, 93

word form, 51, 52

phonologically related words, 82

phonology, 6, 84, 151–158

phrase structure, 164

grammar, 5, 124–127, 130, 133–135, 142,

143, 144, 164, 231

plasticity, 40

plural suffix, 117polysemy, 83–87

positron emission tomography (PET), 44,

46, 60

postsynaptic potential, 19

predicate calculus, 104

prepositions, 115, 126, 188, 214

priming, 6, 7, 83, 170, 172, 173, 174, 177, 182,

184, 200, 201, 202, 205, 217, 218, 223, 229,

230, 231, 242, 244, 245, 252, 258, 265,

266syntactic, 165

principle, of correlation learning, see also

correlation learning, 50, 56, 62, 76,

179

principles, 42, 127

linguistic, 272

neuroscientific, 50, 65, 206, 247, 272

Page 330: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 330/331

Subject Index 313

probabilistic grammars, 106

problems, linguistic, 106

processing indirect sequence, 161, 171

progressive brain diseases, 67projection lines, 141, 231

projections, topographically ordered,

21

projectivity, 141, 143

pronouns, 115, 156, 165, 189, 236, 237,

238, 246, 264

proper names, 189, 225, 226, 231

prototypes, 88

pseudowords, 43, 55, 56, 63–64,

157psycholinguistic theories, 82

psychologists, 33

psychology, 3

psychophysiology, 3

Purkinje cells, 160

pushdown

automaton, 131–133, 262

mechanism, 129, 262

memory, 131, 176, 255, 263

stack, 251storage, 134

store, 244

putamen, 80

R

reading, 46, 95

reciprocal connections, 75, 94, 108, 178,

179

refractory periods, 30

regulation mechanism, 75, 83, 84, 86, 87, 95,175, 194, 205, 246, 266

reorganization, 16

rest, 217, 258

reverberation, 6, 7, 32, 114, 169–174, 177,

180, 182, 183, 185, 200–205, 217, 218,

220–222, 230, 232, 234, 239, 240–242,

252–254, 258, 264, 265, 266

reverberatory

activity, 52–55

synfire chain, 6, 157rhythms, high-frequency, 53

rules, 5, 23, 119, 120, 127, 128, 131, 138–147,

162, 191, 198, 207, 208, 211–213, 238,

239, 273

recursive, 128

rewriting, 125–126, 132

syntactic, 9, 115, 126, 188, 207

S

satisfaction, 185, 216

second language, 91

semanticcategory, 63, 65

dementia, 116, 121

features, 88–89

representations, 86, 88

system, 93, 94

web, 86

semantics, 4, 39, 46, 61–64, 69, 95, 118,

248

sentence processing, 234, 250

sentences, center-embedded, see centerembedding

sequence

detection, 102, 103, 172, 177, 182, 185, 186,

187, 215

detectors, 6, 87, 55, 134, 161–168, 171,

177, 180, 183, 185, 187, 190–198, 216,

222, 226, 227, 231, 235, 253, 257,

266, 267

features, 190, 192, 193, 197, 210

formulas, 7, 209, 210, 217, 226, 227, 257mirror-image, 129, 132

morpheme, 186

sets, 2, 6, 7, 168, 171, 177–204, 208,

210, 215, 218, 220, 221, 223, 227,

230, 251–255, 258, 262, 265, 266

sequencing

sets, see also sequence sets, 168

unit, see also sequence sets, 87,

177, 217

serial order, 23, 107, 128, 144, 146, 147, 151,154, 158–160, 165, 166, 172, 177, 190,

194, 200, 213

information, 269

mechanisms, 107, 159

model, 168

problem, 159

signal-to-noise ratio, 23, 174

signed, 39

silent reading, 44

simulations, 215, 217, 220, 224, 228, 231,232, 262

slash categories, 233

sparse coding, 76–78, 81

sparseness of connections, 94

spatiotemporal

events, 107

patterns, 28, 97, 148, 150, 157, 158, 169

Page 331: Pulv2002NeuroscLan

8/6/2019 Pulv2002NeuroscLan

http://slidepdf.com/reader/full/pulv2002neurosclan 331/331

314 Subject Index

speech

apraxia of, 71

errors, 37, 110

production, 35, 44, 70–71spines, 10, 12

split brain patients, 42, 43

star formation, 270

stochastic models, 107

storage capacity, 112

string detector, 101–103

structure, syntactic, 163, 165, 223, 230

structural

asymmetries, 41

attributes, 60deficits, 67

subordinate clause, 246, 257

subordination, 249

supramarginal gyrus, 38

sylvian fissure, 121

synapses, 10, 12, 13, 18, 76, 102

synaptic strengthening, 22

synchronization, 185, 221, 223, 231, 234,

263

synchrony, 216, 220syndromes of aphasia, see also aphasia,

34

synfire

chains, 6, 147, 148, 149, 150, 151,

152, 153, 155, 156, 158, 161, 171,

240, 271

model, 154, 156, 159

synonymy, 87, 88

syntactic

positive shift, 267processing, 104, 107, 179, 208, 268

representation, 225

theories, 2, 163, 165, 214

tree, 165

violations, 266, 267, 268

syntax, 5, 6, 8, 114, 124, 125, 132, 134, 265,

272

network, 186

theories, 233

T

temporal

areas, 34, 35–36, 46, 50, 52, 69, 72

lobe, see also temporal areas, 39, 41–45,

118

summation, 54

thalamus, 80–81

theoretical neuroscience, 275

theories

linguistic, 147, 238, 271, 272–274

syntactic, 2, 163, 165, 214threshold, 98, 99, 102–104, 148–150, 246

control, 78–79, 175–176, 184, 200–205,

219–222, 221–222, 229–232, 246, 258,

264–265, 269

activation, 102

control mechanism, 176, 183–184, 267

regulation, see threshold control

token test, 34

tools, 58–59

topographical projections, 15–16topographies, 1, 4, 50, 59, 61, 64, 65, 73,

74, 82

traces, 51

transformation, 137, 225, 257

transitional probabilities, 51

translatives, 246, 257, 262, 264

trauma, 33

tree graphs, 141, 143

tumor, 33

Turing machine, 133type-token problem, 145

U

unimodal deficit, 72

universal feature, 95

universal properties of the cortex, 21

V

valence, 95, 139, 187, 209, 249, 256

formulas, 7, 209, 211, 213, 214, 217, 226,227, 250

theory, 138, 214

variable delays, 162

verb, 39, 45, 59, 60, 61, 86, 104–106, 115, 116,

120, 121, 126, 127, 131, 135, 137, 139,

140, 143, 155, 156, 162, 187–192, 197,

208, 210–214, 225, 228, 231, 236, 250,

253, 254, 257

generation, 45

intransitive, 106particle, 106, 134, 137, 138, 140, 188, 189,

211

stem, 105, 119, 121

suffix, 144, 165, 189, 191, 198, 225,

231, 233

visibility, 185, 216, 223, 244

vocabulary, 226, 238